The synaptic cleft manifests enriched glycosylation, with structured glycans coordinating signaling between presynaptic and postsynaptic cells. Glycosylated signaling ligands orchestrating communication are tightly regulated by secreted glycan-binding lectins. Using the Drosophila neuromuscular junction (NMJ) as a model glutamatergic synapse, we identify a new Ca2+-binding (C-type) lectin, Lectin-galC1 (LGC1), which modulates presynaptic function and neurotransmission strength. We find that LGC1 is enriched in motoneuron presynaptic boutons and secreted into the NMJ extracellular synaptomatrix. We show that LGC1 limits locomotor peristalsis and coordinated movement speed, with a specific requirement for synaptic function, but not NMJ architecture. LGC1 controls neurotransmission strength by limiting presynaptic active zone (AZ) and postsynaptic glutamate receptor (GluR) aligned synapse number, reducing both spontaneous and stimulation-evoked synaptic vesicle (SV) release, and capping SV cycling rate. During high-frequency stimulation (HFS), mutants have faster synaptic depression and impaired recovery while replenishing depleted SV pools. Although LGC1 removal increases the number of glutamatergic synapses, we find that LGC1-null mutants exhibit decreased SV density within presynaptic boutons, particularly SV pools at presynaptic active zones. Thus, LGC1 regulates NMJ neurotransmission to modulate coordinated movement.

Synapse assembly and subsequent neurotransmission strength are dependent on secreted bidirectional signaling pathways, which communicate between presynaptic and postsynaptic partners to coordinate synaptic development, function and plasticity (Rushton et al., 2020). Both synaptic cleft and the surrounding extracellular perisynaptic space (the ‘synaptomatrix’) are characterized by heavy glycosylation, which has been shown over the last half-century with extensive lectin imaging (Dani and Broadie, 2012; Dani et al., 2012; Dityatev et al., 2010; Kamimura and Maeda, 2017). Synaptic signals are both mediated by, and regulated via, secreted glycoproteins (GPs), proteoglycans (PGs) and endogenous glycan-binding lectins (Rushton et al., 2020). Disruption of these synaptic glycan-binding and modifying mechanisms is linked to numerous heritable neurological diseases, which include autism spectrum disorder (ASD) and intellectual disability (ID) states such as fragile X syndrome (FXS), movement and degeneration disorders such as hereditary spastic paraplegias (HSPs), and most broadly, the rapidly expanding congenital disorders of glycosylation (CDGs; Doll and Broadie, 2014; Friedman et al., 2013; Orso et al., 2005; Parkinson et al., 2016; Frappaolo et al., 2018). Secreted lectins play important roles in synaptic processes, including the secreted Ca2+-binding (C-type) lectins first identified by Kurt Drickamer. For example, the Drosophila Mind-the-Gap (MTG) patterns synaptomatrix glycans, regulates glutamate receptor (GluR) clustering, and modulates trans-synaptic signaling to control neurotransmission strength (Rohrbough et al., 2007; Rushton et al., 2009, 2012; Rohrbough and Broadie, 2010).

We identified a second C-type lectin, Lectin-galC1 (LGC1), in a whole-genome transgenic Drosophila RNAi screen for genes regulating synapse structure and function at the neuromuscular junction (NMJ; Dani et al., 2012). LGC1 is one of 33 C-type lectins encoded by the Drosophila genome, which are classified by their binding to extracellular β-galactoside glycans via a single characteristic Ca2+-dependent carbohydrate-binding domain (CBD; Ao et al., 2007; Keller and Rademacher, 2019). Similar to the many other lectins encoded by the Drosophila genome, LGC1 exists within a gene cluster with two other duplicated C-type lectin gene homologs (Theopold et al., 1999). However, these two neighbors, with homologies of 59% and 24% respectively, show both differential expression and function (Tanji et al., 2006). Previous LGC1 studies have revealed roles in development and cellular defense, showing enhanced expression in imaginal discs during pupation, binding to surface polysaccharide chains on gram-negative bacteria, and upregulation in response to injury due to an upstream nuclear factor κB-binding site, which is a common regulator of immune defense genes (Haq et al., 1996; Sun and Faye, 1992). Consistent with other glycosylation mechanisms, LGC1 has a pleiotropic role in neurotransmission (Scott and Panin, 2014). Our RNAi screen identification of LGC1 indicated critical roles in the neuromusculature, and specifically at NMJ synapses, with increased neurotransmission strength following LGC1 knockdown (Dani et al., 2012). Our goal here was to pursue this lead by generating selective tools to test LGC1 requirements in coordinated movement and NMJ neurotransmission mechanisms.

In this study, we generated a LGC1 loss-of-function null allele using CRISPR/Cas9 genome editing (Gokcezade et al., 2014) by creating a frame-shift near the beginning of the coding sequence. To characterize neuromuscular roles, we combine this null with a characterized LGC1 deficiency. We generated a LGC1::GFP fusion and specific LGC1 antibody to reveal LGC1 protein at the NMJ, in presynaptic boutons and secreted into the extracellular synaptomatrix. Consistent with this NMJ localization, we found that loss of LGC1 caused faster locomotory muscle peristalsis and much more rapid coordinated movement. Previously, we used LGC1 RNAi to show increased NMJ excitatory junction current (EJC) transmission, but no change in synaptic architecture (Dani et al., 2012). Here, we confirm these preliminary findings in null mutants employing confocal imaging and two-electrode voltage-clamp (TEVC) electrophysiology. Although the loss of LGC1 did not alter NMJ architecture, it did increase synapse number. Consistently, LGC1 mutants showed higher spontaneous synaptic vesicle (SV) fusion rates as well as elevated motor nerve stimulation-evoked EJC amplitudes. To test mechanisms, we employed FM1-43 dye imaging to reveal faster SV cycling dynamics with smaller SV pools in the mutants. With high-frequency stimulation (HFS), we found faster depression and slower recovery. Both confocal and electron microscopy revealed smaller membrane-associated SV pools at mutant active zones, consistent with faster activity-dependent SV cycling. Taken together, these results indicate that LGC1 plays a role in presynaptic SV release dynamics to cap NMJ neurotransmission strength and thus limit motor function.

The LGC1 genetic locus, mutants and NMJ expression

In the Drosophila genome, the C-type lectin gene cluster at cytological location 37D encodes three secreted Ca2+-binding lectins, including Lectin-galC1 (LGC1), CG33532 and CG33533 (Dodd and Drickamer, 2001; Haq et al., 1996). A comparison of the amino acid sequences of these gene products reveals 59% similarity between LGC1 and CG33532, but a lower similarity (24%) between LGC1 and CG33533. All three genes contain two exons, with the second encoding the defining C-type lectin fold (Fig. 1A); but the products have different expression profiles and cell functions (Tanji et al., 2006). The Bloomington Drosophila Stock Center Df(2L)Exel6043 (hereafter designated ‘Df’) completely removes the LGC1 locus, C-type lectin gene cluster and other surrounding genes (Fig. 1A, red bar). To determine specific roles of LGC1 at the NMJ, we created a frame-shift (fs) complete loss-of-function null mutant allele using CRISPR/Cas9 gene editing tools (see Materials and Methods; Fig. 1A, red arrow). This mutation (designated as LGC1fs) generates numerous substitutions and deletions, leading to a small, remnant 70-residue peptide (Fig. 1A, red text). This truncated peptide lacks any known domains, including a C-type lectin carbohydrate-binding domain (CBD). We also made a UAS-LGC1::GFP fusion transgene (hereafter designated ‘L-GFP’), inserted onto chromosome 2 at cytological location 55C4 (Fig. 1B). Both LGC1 mutants and GFP-tagged LGC1 were first used to examine neural and neuromuscular protein expression profiles.

Fig. 1.

LGC1 genomic locus, CRISPR/Cas9 mutation and NMJ expression. (A) Schematic of the LGC1 genomic locus showing exons (blue), untranslated regions (gray) and introns (connecting bars). Df(2L)Exel6043 (red, top) removes the entire LGC1 gene. The CRISPR/Cas9 frame-shift mutation (LGC1fs) is indicated (red arrow). The frame-shift generates a stop codon, resulting in a 70-peptide fragment lacking a carbohydrate-binding domain (CBD). Red text indicates changed nucleotides, and black text indicates conserved nucleotides. (B) Anti-LGC1 western blot of brain lysates from genetic background control (w1118; w), LGC1 homozygous mutant (LGC1fs/LGC1fs; fs), mutant over Df(2L)Exel6043 (LGC1fs/Df; fs/Df), and the transgenic LGC1::GFP fusion line (L-GFP). The anti-LGC1 (25 kDa) band is lost in both LGC1fs/LGC1fs and LGC1fs/Df mutants (bottom, vertical arrows), and LGC1::GFP (52 kDa) is present only in transgenic animals (top, horizontal arrow). (C) Anti-GFP western blot with LGC1::GFP (top, red arrow), and presumed degradation products (bottom arrows). (D) Representative NMJ labeled for the presynaptic membrane marker (HRP, blue) and LGC1 (green) in a w1118 larva. Insets show enlarged bouton labeled for HRP (blue) alone, LGC1 (green) alone, and the merged HRP and LGC1 signals. (E,F) Representative NMJ labeled for presynaptic membrane marker (HRP, blue) and LGC1 (green) in LGC1fs/LGC1fs (E) and LGC1fs/Df (F) larvae. (G) Representative image of 24B-Gal4>LGC1::GFP transgenic fusion in wandering third-instar larva salivary gland (sg), with native fluorescence showing secretion into the extracellular lumen (lum; outlined white dotted lines). (H) Control image of 24B-Gal4>eGFP in wandering third-instar salivary gland, with native fluorescence showing accumulation in sg cytosol and around nuclei. (I) Representative larval NMJ co-labeled for LGC1 (blue), HRP (green) and active zone marker Bruchpilot (BRP, magenta). Inset shows higher magnification image of single bouton co-labeled for LGC1 (blue) and BRP (magenta). (J) Representative larva NMJ co-labeled for LGC1 (white), HRP (green) and postsynaptic scaffold Discs Large (DLG, blue). Inset shows higher magnification single bouton co-labeled for LGC1 (white) and DLG (blue). Images show third-instar larvae and are representative of four separate experiments.

Fig. 1.

LGC1 genomic locus, CRISPR/Cas9 mutation and NMJ expression. (A) Schematic of the LGC1 genomic locus showing exons (blue), untranslated regions (gray) and introns (connecting bars). Df(2L)Exel6043 (red, top) removes the entire LGC1 gene. The CRISPR/Cas9 frame-shift mutation (LGC1fs) is indicated (red arrow). The frame-shift generates a stop codon, resulting in a 70-peptide fragment lacking a carbohydrate-binding domain (CBD). Red text indicates changed nucleotides, and black text indicates conserved nucleotides. (B) Anti-LGC1 western blot of brain lysates from genetic background control (w1118; w), LGC1 homozygous mutant (LGC1fs/LGC1fs; fs), mutant over Df(2L)Exel6043 (LGC1fs/Df; fs/Df), and the transgenic LGC1::GFP fusion line (L-GFP). The anti-LGC1 (25 kDa) band is lost in both LGC1fs/LGC1fs and LGC1fs/Df mutants (bottom, vertical arrows), and LGC1::GFP (52 kDa) is present only in transgenic animals (top, horizontal arrow). (C) Anti-GFP western blot with LGC1::GFP (top, red arrow), and presumed degradation products (bottom arrows). (D) Representative NMJ labeled for the presynaptic membrane marker (HRP, blue) and LGC1 (green) in a w1118 larva. Insets show enlarged bouton labeled for HRP (blue) alone, LGC1 (green) alone, and the merged HRP and LGC1 signals. (E,F) Representative NMJ labeled for presynaptic membrane marker (HRP, blue) and LGC1 (green) in LGC1fs/LGC1fs (E) and LGC1fs/Df (F) larvae. (G) Representative image of 24B-Gal4>LGC1::GFP transgenic fusion in wandering third-instar larva salivary gland (sg), with native fluorescence showing secretion into the extracellular lumen (lum; outlined white dotted lines). (H) Control image of 24B-Gal4>eGFP in wandering third-instar salivary gland, with native fluorescence showing accumulation in sg cytosol and around nuclei. (I) Representative larval NMJ co-labeled for LGC1 (blue), HRP (green) and active zone marker Bruchpilot (BRP, magenta). Inset shows higher magnification image of single bouton co-labeled for LGC1 (blue) and BRP (magenta). (J) Representative larva NMJ co-labeled for LGC1 (white), HRP (green) and postsynaptic scaffold Discs Large (DLG, blue). Inset shows higher magnification single bouton co-labeled for LGC1 (white) and DLG (blue). Images show third-instar larvae and are representative of four separate experiments.

To test LGC1 protein levels, we performed western blots of brain lysates using a new LGC1 antibody (Fig. 1B). Antibodies were generated against the entire sequence, excluding only the signal peptide (see Materials and Methods; Trier et al., 2019). The LGC1 protein has a predicted molecular mass of ∼20 kDa, based on the amino acid sequence prior to post-translational modification. Anti-LGC1 western blots show a clear protein band at ∼24 kDa in the genetic background control (w1118), which is absent in both homozygous LGC1fs/LGC1fs and LGC1fs/Df mutants (Fig. 1B, vertical arrows). This indicates LGC1fs is a protein null, as predicted (Fig. 1A). Ubiquitous daughterless Gal4 (UH1-Gal4; Brand and Perrimon, 1993) driving the UAS-LGC1::GFP transgene (UH1-Gal4>L-GFP) generates a GFP fusion protein with a predicted molecular mass of ∼50 kDa. In addition to the endogenous LGC1 protein band, this line has a unique band at ∼52 kDa (Fig. 1B, horizontal arrow). Anti-GFP western blots confirm the presence of the LGC1::GFP fusion protein, and also confirm the specificity of the anti-LGC1 antibody (Fig. 1C, red arrow). There are also always several smaller GFP+ bands present in the anti-GFP western blots (horizontal arrows), which range in apparent molecular mass from ∼28–38 kDa (Fig. 1C). These bands are assumed to be GFP+ transgene degradation products. These results confirm the LGC1::GFP fusion protein and indicate that LGC1fs is a null mutant allele.

To assess LGC1 neuromuscular expression, we first used the anti-LGC1 antibody with NMJ synaptic marker co-labeling. To test LGC1 secretion, anti-LGC1 antibody (green) was used in detergent-free labeling of larval NMJs, co-labeled with anti-horseradish peroxidase (HRP, blue) to reveal motor neuron membranes (Fig. 1D–F). LGC1 is enriched in NMJ boutons, with additional diffuse labeling surrounding boutons, owing to LGC1 protein secretion (Fig. 1D). LGC1 mutants (LGC1fs/LGC1fs and LGC1fs/Df) do not show neuromuscular labeling or enrichment at the NMJ, consistent with null mutation (Fig. 1E,F). To confirm LGC1 secretion from the well-defined signal peptide, we show 24B-Gal4>L-GFP larvae secrete the LGC1::GFP fusion protein into the salivary gland (sg) lumen (Fig. 1G). Control 24B-Gal4>eGFP larvae show native GFP florescence within sg cells, but no secretion or accumulation in the sg lumen (Fig. 1H). To better localize LGC1 NMJs, animals were co-labeled with the presynaptic active zone marker anti-Bruchpilot (BRP; Fig. 1I; Marrus et al., 2004). LGC1 is enriched in presynaptic boutons (HRP, green) and overlaps with punctate active zones (BRP, magenta), with less LGC1 protein (blue) in NMJ terminal areas lacking synapses (Fig. 1I). Compared with the postsynaptic scaffold marker Discs Large (DLG, Fig. 1J; Thomas et al., 1997), LGC1 (white) is enriched in and adjacent to synaptic boutons, well within DLG-marked (blue) subsynaptic reticulum (Fig. 1J). These results show LGC1 is in the presynaptic NMJ boutons, secreted, present in the extracellular space immediately around boutons, and fully contained within the encircling postsynaptic domain.

LGC1 regulates locomotory peristalsis and coordinated movement

Since NMJs mediate neurotransmission from motor neurons to muscles to drive muscle contraction, we hypothesized that LGC1 could influence coordinated movement. We previously identified LGC1 as a negative regulator of NMJ neurotransmission based on a genomic transgenic RNAi screen (Dani et al., 2012). LGC1 RNAi knockdown causes a significant increase in motoneuron stimulation-evoked NMJ transmission, but it was not at all clear how this might translate to changes in the overall animal movement function. Muscle contraction can be accurately quantified (Chandel et al., 2019). Drosophila larval body segments contain a defined motor circuit and muscle pattern, with contraction and relaxation generating peristaltic waves that propel the larva through the environment (Fushiki et al., 2016; Gjorgjieva et al., 2013). NMJ mechanisms underlie the sequential temporal and spatial activation of discrete muscles in larvae that mediate coordinated movement (Grillner, 2003). Locomotory peristaltic waves are one measurement of basal motor function driving normal movement (Nichols et al., 2012; Pulver et al., 2015). A simple roll-over assay that requires sequential muscle control to right a larva from an inverted position measures more complex, coordinated movement (Bodily et al., 2001; Jumbo-Lucioni et al., 2016). We use both approaches to compare genetic background control (w1118), homozygous LGC1fs/LGC1fs and LGC1fs/Df mutants. Representative still image frames from locomotion and rollover videos are shown in Fig. 2.

Fig. 2.

LGC1 limits locomotory peristaltic waves and coordinated movement. (A) Still frames of movies showing wandering third-instar peristaltic muscle waves in the genetic background control (w1118). Arrows indicate position of wave propagation from posterior initiation to anterior termination. (B) Quantification of peristaltic wave duration in the genetic control (w1118), the LGC1-null mutant (LGC1fs) and the LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df). Scatterplots show all data points with mean±s.e.m. Sample size, n=15 animals/genotype, with n=5 peristaltic waves/animal. ***P<0.001 (Welch's t-test). (C) Still frames of movies showing larval rollover movement in the genetic background control (w1118). Dotted line indicates dorsal midline towards agarose plate, solid line indicates midline upwards from plate. (D) Quantification of rollover duration measuring coordinated movement reaction time in the w1118 control (w1118), LGC1 null (LGC1fs) and Df trans-heterozygote (LGC1fs/Df). Scatterplots show all data points with mean±s.e.m. Sample size: n=10 animals/genotype, with n=3 rollover assays/animal for each point. ****P<0.0001 (Welch's t-test).

Fig. 2.

LGC1 limits locomotory peristaltic waves and coordinated movement. (A) Still frames of movies showing wandering third-instar peristaltic muscle waves in the genetic background control (w1118). Arrows indicate position of wave propagation from posterior initiation to anterior termination. (B) Quantification of peristaltic wave duration in the genetic control (w1118), the LGC1-null mutant (LGC1fs) and the LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df). Scatterplots show all data points with mean±s.e.m. Sample size, n=15 animals/genotype, with n=5 peristaltic waves/animal. ***P<0.001 (Welch's t-test). (C) Still frames of movies showing larval rollover movement in the genetic background control (w1118). Dotted line indicates dorsal midline towards agarose plate, solid line indicates midline upwards from plate. (D) Quantification of rollover duration measuring coordinated movement reaction time in the w1118 control (w1118), LGC1 null (LGC1fs) and Df trans-heterozygote (LGC1fs/Df). Scatterplots show all data points with mean±s.e.m. Sample size: n=10 animals/genotype, with n=3 rollover assays/animal for each point. ****P<0.0001 (Welch's t-test).

Movement peristalsis was measured in free-moving wandering third-instar larvae during the linear phase of locomotion, with the contraction wave proceeding sequentially along the entire body length (Fig. 2A, arrows). In w1118 controls, this peristaltic wave takes just over 1 s to travel the body length (Fig. 2B). In LGC1 mutants (LGC1fs and LGC1fs/Df), the waves proceed in a significantly (P<0.001) more rapid segmental sequence to drive quicker forward movement, compared to slower peristaltic waves in matched controls (Fig. 2B). Coordinated movement was measured using the rollover assay on manipulated third-instar larvae, which requires a rapid sequence of bilaterally coordinated timed muscle contractions to allow the larvae to turn from an inverted position to a dorsal upwards position (Fig. 2C). In w1118 controls, this righting behavior takes 9.98±1.00 s (mean±s.e.m.) on average (Fig. 2D). In contrast, LGC1fs and LGC1fs/Df null mutants exhibit much quicker coordinated movement that enables a very rapid righting behavior (LGC1fs, 2.92±0.20 s; LGC1fs/Df, 4.84±0.37 s; P<0.0001, n=30; Fig. 2D). These results indicate that LGC1 limits both peristaltic muscle waves during basal locomotion and coordinated movement during complex motor behaviors. Given LGC1 localization at the NMJ, these phenotypes suggest changes in NMJ muscle control in LGC1 mutants, owing to alterations in NMJ connectivity/architecture, synapse number and/or transmission properties from altered synaptic signaling properties.

LGC1 limits glutamatergic synapse number at the NMJ

The presence of the very significant behavioral phenotype of highly accelerated coordinated movement led us to hypothesize possible increased synapse structure or function at the NMJ. Over-elaboration in the number of NMJ synaptic boutons has been causally associated with more rapid movement (Jumbo-Lucioni et al., 2014, 2016). Likewise, increased NMJ synapse number (presynaptic active zones and postsynaptic glutamate receptors) elevates neurotransmission to drive faster muscular movement (Kittel et al., 2006; Rasse et al., 2005; Wagh et al., 2006). To test these possibilities, we assayed for changes in NMJ structure and synapse number in LGC1 mutants versus matched controls. To test structure, NMJs were co-labeled for presynaptic anti-HRP and postsynaptic anti-DLG (Fig. 3). In w1118 controls, NMJs show a consistent architecture of regularly spaced synaptic boutons (Fig. 3A). In the LGC1-null mutants (LGC1fs and LGC1fs/Df), there is no detectable change in this usual NMJ architecture (Fig. 3B,C). In quantified measurements, there is no significant difference in NMJ area, branching or synaptic bouton number (Fig. 3D–F). These findings are consistent with our previous RNAi results, which also show no changes in NMJ area, branching or synaptic bouton number with LGC1 RNAi knockdown (Dani et al., 2012). Therefore, LGC1 has no detectable role in synaptic morphological development, but it remained possible that LGC1 may have a role in regulating synapse number within the normal NMJ boutons.

Fig. 3.

LGC1 loss does not affect NMJ structure, but elevates synapse density. (A) Representative confocal image of wandering third-instar NMJ (muscle 4) co-labeled for presynaptic HRP (blue) and postsynaptic DLG (magenta) in the genetic background control (w1118). N indicates nerve bundles; arrows point to NMJ boutons. (B) LGC1 homozygous null mutant (LGC1fs) labeled as above. (C) LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df) labeled as above. (D) Quantification of synaptic bouton number. (E) Quantification of NMJ axonal branch number. (F) Quantification of NMJ area. (G–I) The same genotypes co-labeled for synaptic active zone BRP (green) and the glutamate receptor II subunit C (GluRIIC, red) at NMJs. Arrows indicate representative boutons. (J–L) High-magnification images of single NMJ boutons. Crosses (+) mark BRP puncta. (M) Quantification of synapse density per NMJ. Scatterplots show all data points with mean±s.e.m. Sample size, n=10 (B) and n=12 (E) animals/genotype. *P<0.05; ns, not significant (Welch's t-test).

Fig. 3.

LGC1 loss does not affect NMJ structure, but elevates synapse density. (A) Representative confocal image of wandering third-instar NMJ (muscle 4) co-labeled for presynaptic HRP (blue) and postsynaptic DLG (magenta) in the genetic background control (w1118). N indicates nerve bundles; arrows point to NMJ boutons. (B) LGC1 homozygous null mutant (LGC1fs) labeled as above. (C) LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df) labeled as above. (D) Quantification of synaptic bouton number. (E) Quantification of NMJ axonal branch number. (F) Quantification of NMJ area. (G–I) The same genotypes co-labeled for synaptic active zone BRP (green) and the glutamate receptor II subunit C (GluRIIC, red) at NMJs. Arrows indicate representative boutons. (J–L) High-magnification images of single NMJ boutons. Crosses (+) mark BRP puncta. (M) Quantification of synapse density per NMJ. Scatterplots show all data points with mean±s.e.m. Sample size, n=10 (B) and n=12 (E) animals/genotype. *P<0.05; ns, not significant (Welch's t-test).

To address this possibility, we assayed LGC1-null mutants (LGC1fs and LGC1fs/Df) for changes in NMJ synapse number. Wandering third-instar larvae were double-labeled for presynaptic and postsynaptic components. The presynaptic active zone (AZ) marker anti-Bruchpilot (BRP) was used to mark synaptic vesicle (SV) release sites in NMJ boutons (Fig. 3G–L; Wagh et al., 2006). Postsynaptic glutamate receptor (GluR) domains were marked with an antibody against a subunit present in all these receptors (anti-GluRIIC) to label clusters in the muscle membrane (Fig. 3G–L; Marrus and DiAntonio, 2004). In w1118 controls, each presynaptic AZ directly apposes a postsynaptic GluR cluster to mediate fast NMJ neurotransmission (Fig. 3G,J; Schuster et al., 1991). Synapse density, measured as synapse number per bouton area is 1.60±0.08/µm2 (mean±s.e.m.; Fig. 3M). In comparison, LGC1 loss caused an increase in synapse density, with an elevated number of BRP and GluRIIC puncta (Fig. 3H,I,K,L). Both homozygous null LGC1fs/LGC1fs and the LGC1fs/Df exhibit a small but significant (P<0.05) increase in synapse number per NMJ bouton area (LGC1fs, 1.81±0.05/µm2; n=12, P=0.03; LGC1fs/Df, 1.87±0.06/µm2; n=12, P=0.01; Fig. 3M). This increase alone is unlikely to explain the much more significant elevation in coordinated movement in LGC1 mutants. Moreover, although structural and functional differentiation can be related, they are independent processes (Menon et al., 2013). Therefore, we next assayed whether LGC1 plays a role in NMJ neurotransmission function.

LGC1 negatively regulates SV release and neurotransmission strength

Our RNAi screen indicated that LGC1 knockdown elevates neurotransmission strength (Dani et al., 2012). However, RNAi knockdown only causes a partial loss of function, and there is always the concern of possible off-target RNAi effects (Seinen et al., 2011; Heigwer et al., 2018). We therefore next tested the contributions of LGC1 to NMJ function using electrophysiological recordings in our new CRISPR/Cas9-generated null mutant alone and in trans to deficiency. A two-electrode voltage-clamp (TEVC) configuration was used to make both motor nerve stimulation-evoked excitatory junction current (EJC) and spontaneous miniature EJC (mEJC) recordings using 1.0 mM [Ca2+] physiological saline (Parkinson et al., 2016; Kopke et al., 2020). For mEJC recordings, the motor nerve was severed and the muscle was voltage-clamped at −60 mV, with continuous records made in 2-min sessions. For evoked EJCs, a suction electrode was used to stimulate the motor nerve innervating the muscle at a suprathreshold voltage (Rohrbough et al., 1999). Repeated recordings (0.2 Hz, 10 consecutive stimuli) were averaged to determine the mean peak EJC amplitude for each animal (N=1). TEVC records were made from staged wandering third-instar larvae comparing the genetic background control (w1118) to the homozygous LGC1fs/LGC1fs null mutants and LGC1fs/Df. Representative evoked EJC and spontaneous mEJC traces, as well as their quantified measurements, are all shown in Fig. 4.

Fig. 4.

Elevated synaptic vesicle fusion and transmission in LGC1-null NMJs. Electrophysiological recordings from wandering third-instar larva NMJs made in two-electrode voltage-clamp (TEVC) configuration. (A) Representative motor nerve stimulation-evoked excitatory junction current (EJC) traces (1.0 mM Ca2+) from genetic background control (w1118, left), an LGC1-null mutant (LGC1fs, middle) and the LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, right). (B) Quantification of EJC mean peak amplitudes. Scatterplots show all data points with the mean±s.e.m. Sample size, n≥25 animals per genotype. *P≤0.05 (Welch's t-test). (C) Representative miniature EJC (mEJC) traces (1.0 mM Ca2+) from the above three genotypes. Quantification of mEJC frequency (D) and mEJC amplitude (E). Scatterplots show all data points with the mean±s.e.m. Sample size, n≥10 animals per genotype. ***P<0.001; ns, not significant (Welch's t-test).

Fig. 4.

Elevated synaptic vesicle fusion and transmission in LGC1-null NMJs. Electrophysiological recordings from wandering third-instar larva NMJs made in two-electrode voltage-clamp (TEVC) configuration. (A) Representative motor nerve stimulation-evoked excitatory junction current (EJC) traces (1.0 mM Ca2+) from genetic background control (w1118, left), an LGC1-null mutant (LGC1fs, middle) and the LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, right). (B) Quantification of EJC mean peak amplitudes. Scatterplots show all data points with the mean±s.e.m. Sample size, n≥25 animals per genotype. *P≤0.05 (Welch's t-test). (C) Representative miniature EJC (mEJC) traces (1.0 mM Ca2+) from the above three genotypes. Quantification of mEJC frequency (D) and mEJC amplitude (E). Scatterplots show all data points with the mean±s.e.m. Sample size, n≥10 animals per genotype. ***P<0.001; ns, not significant (Welch's t-test).

Nerve stimulation results in robust NMJ glutamatergic transmission. In the w1118 controls, muscle TEVC records show a reliable, high fidelity EJC following each motor nerve stimulation (Fig. 4A). The control EJC amplitude averaged from 10 consecutive recordings is 154.8±7.4 nA (mean±s.e.m.; Fig. 4B). Both LGC1-null mutants (LGC1fs/LGC1fs and LGC1fs/Df) exhibited clearly larger EJC responses (Fig. 4A, right), consistent with our earlier LGC1 RNAi results (Dani et al., 2012). Quantification reveals similar evoked EJC transmission amplitudes of 184.8±8.8 nA for LGC1fs and 176.3±7.2 nA for LGC1fs/Df (Fig. 4B). The amplitudes were not significantly different from each other, but both LGC1 nulls were significantly increased compared with genetic background controls (Fig. 4B). The mutant phenotype reflects a selective elevation in the peak EJC neurotransmission amplitude without any detectable changes in EJC kinetics, including stimulation latency, EJC rise time and decay constant (Fig. 4A). An increased EJC indicates a strengthened NMJ due to either greater quantal content, with more SVs released during evoked nerve stimulation (Pan and Zucker, 2009), or a greater number/function of the postsynaptic glutamate receptors (Astorga et al., 2016; Gerkin et al., 2013), or both together. To distinguish these possibilities, we next assayed spontaneous mEJCs, to test presynaptic vesicle fusion probability (mEJC frequency), as well as the number of activated postsynaptic receptors (mEJC amplitude).

We used mEJC recordings to begin to dissect the presynaptic and postsynaptic mechanisms regulated by LGC1 (Bykhovskaia and Vasin, 2017). Representative mEJC traces from all three genotypes are shown in Fig. 4C. In the w1118 controls, mEJC events occur at a regular low frequency and amplitude (Fig. 4C, top). Quantification reveals an average control mEJC frequency of 0.50±0.07 Hz, with an average control mEJC amplitude of 0.32±0.06 nA (mean±s.e.m.; Fig. 4D,E). In contrast, both LGC1-null mutants (LGC1fs/LGC1fs and LGC1fs/Df) exhibit an obvious increase in mEJC frequency, albeit without any apparent change in the mEJC amplitude (Fig. 4C, middle and bottom). Quantification reveals that mEJC frequency is significantly higher (P<0.001) in both LGC1fs (1.94±0.30 Hz) and LGC1fs/Df (2.07±0.29 Hz), without any change in either of the mutant mEJC amplitudes (Fig. 4D,E). These results indicate that LGC1 limits evoked neurotransmission and spontaneous vesicle release. Therefore, the LGC1-null effects are primarily presynaptic, with an increase in SV fusion probability that elevates evoked quantal content in the absence of LGC1 function. The increased evoked EJC amplitude in LGC1-null mutants perhaps reflects the greater synapse number (Fig. 3J–M), but also likely increased SV release probability at each synapse (Fig. 4A–D). Moreover, the functional strengthening suggests additional effects on SV dynamics, which we therefore next analyzed using live optical imaging.

LGC1 increases SV cycling rate at the NMJ

The above findings indicate that LGC1 function limits neurotransmission strength, due to an elevated presynaptic fusion probability. Since SV cycling dynamics drive presynaptic vesicle availability and therefore their release capabilities (Gan and Watanabe, 2018; Chanaday et al., 2019), we next tested LGC1 roles in the SV cycle underlying functional neurotransmission. We employed the activity-dependent, depolarization-driven lipophilic FM1-43 dye labeling technique to monitor the SV cycle (Betz and Bewick, 1992), which is widely used to quantitatively measure SV endocytosis and exocytosis at the Drosophila NMJ (Kopke and Broadie, 2018). In this approach, the wandering third-instar larva neuromusculature is bathed in FM1-43 dye (4 µM), which is incorporated into cycling SVs (‘loaded’) through endocytosis after stimulation with a motor nerve suction electrode as above (Kopke et al., 2017). This loading provides a quantitative measurement of the functional SV cycling pool size (Staples and Broadie, 2013; Vijayakrishnan et al., 2009). A second electrical stimulation of the motor nerve causes depolarization in the absence of FM1-43 dye, driving SV fusion and release of the FM1-43 dye (‘unloaded’) through exocytosis. This provides a measurement of SV release efficacy from presynaptic active zones. Representative low- and high-magnification images comparing FM1-43 dye loading and unloading in genetic background control (w1118), and the LGC1fs and LGC1fs/Df null mutants are shown in Fig. 5.

Fig. 5.

LGC1 loss reduces synaptic vesicle pool size and increases cycling rate. (A–F) Representative confocal images of wandering third-instar larva NMJs (muscle 4) labeled for anti-HRP (blue) and with FM1-43 intensity shown as a heat-map (with indicated scale) in the genetic background control (w1118, top), LGC1 null (LGC1fs, middle) and LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, bottom). A, C and E show the FM1-43 intensity after 20 Hz nerve stimulation for 5 min (load). B, D and F show FM1-43 intensity after 20 Hz nerve stimulation in the absence of dye (unload). Arrows indicate representative NMJ boutons. (G–L) High-magnification images of single boutons in the same three genotypes, with FM1-43 loaded images in G, I and K, and after unloading in H, J and L. (M) Quantification of synaptic vesicle FM1-43 dye loading and unloading in the above three genotypes. Scatterplots show all data points with the mean±s.e.m. Circles show loaded intensities and squares show unloaded intensities. Sample size, n≥10 animals per genotype. *P≤0.01; **P≤0.001 (Dunn's multiple comparisons test).

Fig. 5.

LGC1 loss reduces synaptic vesicle pool size and increases cycling rate. (A–F) Representative confocal images of wandering third-instar larva NMJs (muscle 4) labeled for anti-HRP (blue) and with FM1-43 intensity shown as a heat-map (with indicated scale) in the genetic background control (w1118, top), LGC1 null (LGC1fs, middle) and LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, bottom). A, C and E show the FM1-43 intensity after 20 Hz nerve stimulation for 5 min (load). B, D and F show FM1-43 intensity after 20 Hz nerve stimulation in the absence of dye (unload). Arrows indicate representative NMJ boutons. (G–L) High-magnification images of single boutons in the same three genotypes, with FM1-43 loaded images in G, I and K, and after unloading in H, J and L. (M) Quantification of synaptic vesicle FM1-43 dye loading and unloading in the above three genotypes. Scatterplots show all data points with the mean±s.e.m. Circles show loaded intensities and squares show unloaded intensities. Sample size, n≥10 animals per genotype. *P≤0.01; **P≤0.001 (Dunn's multiple comparisons test).

In w1118 control NMJs, boutons load FM1-43 dye upon motor nerve stimulation, with boutons loaded strongly throughout the synaptic terminal (Fig. 5A) and intense fluorescence within individual presynaptic boutons (Fig. 5G). In LGC1-null mutants, there is significantly less dye incorporation throughout the entire NMJ (Fig. 5C,E) and within individual synaptic boutons (Fig. 5I,K). Compared to the control FM1-43 uptake (56.6±4.9, mean±s.e.m.), both LGC1fs (32.1±6.6) and LGC1fs/Df (24.0±3.1) exhibit very significantly less dye loading (Fig. 5M). The w1118 control NMJs unload most of the FM1-43 upon the second stimulation (Fig. 5B), although some of the dye is still retained within individual NMJ boutons (Fig. 5H). In contrast, the LGC1-null mutants show negligible amounts of dye remaining after unloading (Fig. 5D,F). Importantly, both LGC1-null boutons retain much less FM1-43 than control boutons following unloading (Fig. 5J,L). The quantified measurements show LGC1 nulls display significantly (P<0.01) elevated SV exocytosis, releasing more dye from their presynaptic boutons (w1118 control, 22.8±2.8; LGC1fs, 9.2±2.3; LGC1fs/Df, 7.5±1.8; Fig. 5M). Thus, both LGC1fs and LGC1fs/Df mutants appear to have smaller SV cycling pools, with much less of the FM1-43 dye taken up, but a faster rate of SV cycling, with higher levels of exocytosis. These results suggest a likely increase in synaptic depression during high frequency usage owing to a smaller SV pool and faster activity-dependent SV cycling.

LGC1 impairs activity-dependent SV cycling at the NMJ

The mismatch in SV pool size versus release rate suggests LGC1-null NMJs will fatigue more rapidly (Fergestad et al., 1999; Rudling et al., 2018). At the Drosophila NMJ, high-frequency neurotransmission challenges the synaptic release machinery, especially at high-probability release sites, which exhibit activity-dependent depression (Kauwe and Isacoff, 2013). Following high activity levels, new SV generation and SV recruitment enables recovery of neurotransmission amplitudes within tens of seconds (Bui and Glavinović, 2014). We therefore hypothesized that the LGC1 mutant SV cycling defects determined with FM1-43 imaging (Fig. 5) would cause faster activity-dependent synaptic depression as well as slower recovery from this depression. To test this idea, variable high frequency stimulation (HFS; 5, 20 and 50 Hz; Stevens et al., 2012; Kauwe and Isacoff, 2013; Newman et al., 2017) in conjunction with TEVC recording was used to measure both synaptic depression and recovery (Fig. 6A). The HFS paradigm used low frequency (0.2 Hz) stimulation to establish the EJC amplitude baseline, following by HFS for 5 min (5, 20 or 50 Hz) to measure EJC amplitude depression. In addition to studying depression, the following EJC amplitude recovery was measured at 30 s intervals at basal 0.2 Hz stimulation (Fig. 6A). Sample TEVC traces of the genetic background control (w1118) compared to LGC1fs and LGC1fs/Df mutants, together with detailed frequency quantifications, are all shown in Fig. 6.

Fig. 6.

Elevated transmission depression and reduced recovery in LGC1 nulls. Electrophysiological recordings from wandering third-instar larva NMJs made in two-electrode voltage-clamp (TEVC) configuration. (A) Schematic of high frequency stimulation (HFS) paradigm at 5 Hz, 20 Hz or 50 Hz for 5 min (depression), followed by 10-train 0.2 Hz recordings every 30 s (recovery). (B) Representative nerve stimulation-evoked excitatory junction current (EJC) traces (1.0 mM Ca2+) from genetic background control (w1118, left), LGC1 null (LGC1fs, middle) and LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, bottom). EJCs shown at t=0 s, black; t=1 s, orange; t=2 min, blue; t=5 min, purple; t=+30 s during recovery, red. (C) Quantification of EJC amplitude over time normalized to initial EJC amplitude for each genotype. HFS train on left (depression, white background) and recovery period on the right (red background). (D) Quantification of EJC amplitudes from indicated time points in C. Scatterplots show all the data points with mean±s.e.m. Sample size, n≥7 animals/genotype. *P<0.05; **P<0.01; ***P<0.001 (one-way ANOVA with post-hoc Kruskal–Wallis tests).

Fig. 6.

Elevated transmission depression and reduced recovery in LGC1 nulls. Electrophysiological recordings from wandering third-instar larva NMJs made in two-electrode voltage-clamp (TEVC) configuration. (A) Schematic of high frequency stimulation (HFS) paradigm at 5 Hz, 20 Hz or 50 Hz for 5 min (depression), followed by 10-train 0.2 Hz recordings every 30 s (recovery). (B) Representative nerve stimulation-evoked excitatory junction current (EJC) traces (1.0 mM Ca2+) from genetic background control (w1118, left), LGC1 null (LGC1fs, middle) and LGC1 null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df, bottom). EJCs shown at t=0 s, black; t=1 s, orange; t=2 min, blue; t=5 min, purple; t=+30 s during recovery, red. (C) Quantification of EJC amplitude over time normalized to initial EJC amplitude for each genotype. HFS train on left (depression, white background) and recovery period on the right (red background). (D) Quantification of EJC amplitudes from indicated time points in C. Scatterplots show all the data points with mean±s.e.m. Sample size, n≥7 animals/genotype. *P<0.05; **P<0.01; ***P<0.001 (one-way ANOVA with post-hoc Kruskal–Wallis tests).

Representative EJC recordings from all three genotypes at specified HFS time points (t=0, 1 s, 2 min and 5 min) reveal the relative depression of synaptic transmission over time (Fig. 6B). After 30 s of low frequency rest, the fifth trace (+30 s) shows the relative partial recovery of EJC amplitude in the different genotypes (Fig. 6B). The control and LGC1 nulls all show progressive presynaptic depression with increasing HFS frequencies (5–50 Hz; Fig. 6C). However, LGC1fs and LGC1fs/Df show significantly decreased EJC amplitudes immediately following the 5 Hz HFS train initiation (Fig. 6C). LGC1 loss causes greater SV release in <1 s with HFS, consistent with the greater SV release probability in nulls (Fig. 6D). At higher frequencies (20–50 Hz), significantly greater synaptic depression no longer occurs in LGC1 nulls, presumably due to heightened activity-dependent demand depressing controls beyond tolerance (Fig. 6C). Controls and LGC1 nulls all show progressive reduced recovery with increasing HFS frequencies (5–50 Hz; Fig. 6C, right, red shading). However, both LGC1fs and LGC1fs/Df have significantly slower recovery rates. Following all HFS trains (5–50 Hz), LGC1 nulls show reduced recovery at 30 s normalized to the starting basal EJC amplitudes (5 Hz, w1118, 0.88±0.03; LGC1fs, 0.53±0.08; 20 Hz, w1118, 0.69±0.05; LGC1fs, 0.41±0.04; and 50 Hz, w1118, 0.59±0.07; LGC1fs, 0.27±0.07; Fig. 6D). These functional defects suggest LGC1 loss alters presynaptic SV availability, which we next attempted to image.

LGC1 regulates SV membrane association at active zones

In order to visualize presynaptic SV distribution in control boutons compared to LGC1-null mutants, we combined confocal imaging for SV pools (Daniels et al., 2004; Martin and Krantz, 2014) with imaging via transmission electron microscopy (TEM) ultrastructure studies to directly image individual SVs (Dear et al., 2016; Kopke and Broadie, 2018). We first used immunocytochemistry to label NMJ synaptic boutons with anti-vesicular glutamate transporter (VGLUT), an integral vesicle membrane protein that specifically labels all glutamatergic SVs (Wilson et al., 2005). Confocal imaging of this marker provides a quantitative measurement of SV number within entire NMJ boutons, as well as some means to map SV distribution relative to the presynaptic membrane in controls compared to LGC1fs and LGC1fs/Df (Fig. 7). TEM studies reveal the synaptic bouton ultrastructure, including the number and distribution of presynaptic active zones, SV density relative to active zones, and the muscle membrane subsynaptic reticulum (SSR) architecture (Wagner, 2017). We quantified all of these synaptic ultrastructural features in genetic background control (w1118) compared to LGC1fs null mutants (Fig. 8). However, the focus of both the light and electron microscopy imaging was on SV density and SV distribution within individual NMJ synaptic boutons, to provide better mechanistic insights into the movement behavior and functional NMJ neurotransmission changes characterizing the LGC1-null condition.

Fig. 7.

LGC1 loss reduces synaptic vesicle density and membrane association. (A) Representative confocal image of a wandering third-instar larva NMJ (muscle 4) co-labeled for presynaptic membrane (HRP, blue) and the synaptic vesicle glutamate transporter (VGLUT, white) in genetic background control (w1118). N indicates nerve bundles; arrows point to NMJ boutons. Insets show VGLUT alone (top), HRP alone (middle) and the merge (bottom). Note dense cortical labeling of VGLUT-marked SVs. (B) An LGC1-null (LGC1fs) larva NMJ labeled as above. Note reduction of cortical VGLUT labeling in the mutant inset. (C) An LGC1-null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df) larva NMJ labeled as above, showing the same phenotype. (D) Quantification of anti-VGLUT fluorescence intensity in the above three genotypes. Scatterplots show all data points with mean±s.e.m. Sample size, n=18 animals/genotype. *P<0.05 (Welch's t-test). (E–G) High-magnification confocal images of single boutons from the above three genotypes, labeled for VGLUT in synaptic vesicles. The yellow bar indicates the line-scan orientation. (H–J) Quantification of line-scans in the above three genotypes. Gray values indicate VGLUT fluorescence intensity, and the x-axis shows distance (µm) across the synaptic bouton. Note the genetic background control (w1118) shows that plasma membrane associated intensity peaks are lacking in both of the LGC1-null conditions (H).

Fig. 7.

LGC1 loss reduces synaptic vesicle density and membrane association. (A) Representative confocal image of a wandering third-instar larva NMJ (muscle 4) co-labeled for presynaptic membrane (HRP, blue) and the synaptic vesicle glutamate transporter (VGLUT, white) in genetic background control (w1118). N indicates nerve bundles; arrows point to NMJ boutons. Insets show VGLUT alone (top), HRP alone (middle) and the merge (bottom). Note dense cortical labeling of VGLUT-marked SVs. (B) An LGC1-null (LGC1fs) larva NMJ labeled as above. Note reduction of cortical VGLUT labeling in the mutant inset. (C) An LGC1-null trans-heterozygous over Df(2L)Exel6043 (LGC1fs/Df) larva NMJ labeled as above, showing the same phenotype. (D) Quantification of anti-VGLUT fluorescence intensity in the above three genotypes. Scatterplots show all data points with mean±s.e.m. Sample size, n=18 animals/genotype. *P<0.05 (Welch's t-test). (E–G) High-magnification confocal images of single boutons from the above three genotypes, labeled for VGLUT in synaptic vesicles. The yellow bar indicates the line-scan orientation. (H–J) Quantification of line-scans in the above three genotypes. Gray values indicate VGLUT fluorescence intensity, and the x-axis shows distance (µm) across the synaptic bouton. Note the genetic background control (w1118) shows that plasma membrane associated intensity peaks are lacking in both of the LGC1-null conditions (H).

Fig. 8.

LGC1 loss reduces synaptic vesicle density surrounding active zones. (A) A representative low-magnification TEM ultrastructure image of a NMJ bouton in the genetic background control (w1118). The profile shows presynaptic active zones (arrows) surrounded by synaptic vesicle pools and muscle subsynaptic reticulum (SSR) membrane folds. (B) High-magnification image of a single active zone (AZ) in the genetic background control (w1118), showing synaptic vesicles surrounding the electron-dense t-bar. Yellow lines show zones used to quantify the synaptic vesicle distribution, with diameters of 100 nm, 200 nm and 300 nm from the active zone t-bar. (C) Low-magnification transmission electron microscope (TEM) image of a NMJ bouton in the LGC1-null mutant (LGC1fs). (D) High-magnification image of a single active zone (AZ) in the LGC1-null mutant (LGC1fs). (E) Quantification of mean synaptic vesicle density within the above three zones comparing controls and null mutants. Sample size, n≥18 animals/genotype. ****P≤0.0001 (one-way ANOVA with post-hoc Mann–Whitney tests). (F) Quantification of all synaptic vesicle densities within each of the above three zones. Scatterplots show all data points with mean±s.e.m. *P≤0.05; **P≤0.01 (two-way ANOVA with post-hoc Mann–Whitney tests).

Fig. 8.

LGC1 loss reduces synaptic vesicle density surrounding active zones. (A) A representative low-magnification TEM ultrastructure image of a NMJ bouton in the genetic background control (w1118). The profile shows presynaptic active zones (arrows) surrounded by synaptic vesicle pools and muscle subsynaptic reticulum (SSR) membrane folds. (B) High-magnification image of a single active zone (AZ) in the genetic background control (w1118), showing synaptic vesicles surrounding the electron-dense t-bar. Yellow lines show zones used to quantify the synaptic vesicle distribution, with diameters of 100 nm, 200 nm and 300 nm from the active zone t-bar. (C) Low-magnification transmission electron microscope (TEM) image of a NMJ bouton in the LGC1-null mutant (LGC1fs). (D) High-magnification image of a single active zone (AZ) in the LGC1-null mutant (LGC1fs). (E) Quantification of mean synaptic vesicle density within the above three zones comparing controls and null mutants. Sample size, n≥18 animals/genotype. ****P≤0.0001 (one-way ANOVA with post-hoc Mann–Whitney tests). (F) Quantification of all synaptic vesicle densities within each of the above three zones. Scatterplots show all data points with mean±s.e.m. *P≤0.05; **P≤0.01 (two-way ANOVA with post-hoc Mann–Whitney tests).

In w1118 control NMJs, VGLUT-marked SVs are restricted to synaptic boutons in a dense cortical ring immediately underlying the plasma membrane (Fig. 7A, arrow inset). In LGC1 nulls, this labeling is clearly and consistently reduced (Fig. 7B,C). Compared to the VGLUT intensity in controls (49.6±2.5, mean±s.e.m.), the mean intensity is significantly (P<0.05) decreased in both the LGC1fs (40.6±2.2) and LGC1fs/Df (41.8±2.4) mutants (Fig. 7D). This signal loss suggests a lower SV density within NMJ boutons, which we confirmed with TEM imaging (Fig. 8). To examine the SV distribution within individual boutons, higher magnification VGLUT distribution measurements were compared between genotypes. LGC1 nulls show a relatively uniform, lower intensity fluorescence within NMJ boutons, while controls have increased VGLUT near the presynaptic membrane (Fig. 7E–G). In w1118 controls, line graphs of fluorescence intensity show a clear decrease in signal in the bouton interior with peaks associated with each membrane (Fig. 7E,H). In sharp contrast, both the LGC1fs and LGC1fs/Df null mutants show a relatively uniform VGLUT signal distribution across the entire synaptic bouton, and lack the two membrane peaks (Fig. 7F,G,I,J). Note that the lowest fluorescence values for all three genotypes are similar, suggesting that mutants are selectively deficient in SVs near the membrane. These results indicate altered SV distributions in mutants at the whole-bouton level, but this would be better supported by SV imaging at single vesicle resolution.

TEM studies show almost no synaptic ultrastructural changes in LGC1 nulls, with normal synaptic boutons, SVs, AZs and postsynaptic SSR compared to w1118 controls (Fig. 8A–D). To examine SV distribution, numbers were quantified within three consecutive 100-nm diameter semicircular zones at AZs (Fig. 8B,D, yellow zones 1–3). Studies show a decrease in SV density across all zones in the LGC1 mutants (Fig. 8D). Quantification shows both controls and the LGC1fs null exhibit a decrease in SV density from zone 1 to 3, with mutants showing lower density in all zones. The SV density decrease in zone 1 and 2 is particularly significant in LGC1fs nulls (P<0.0001; Fig. 8E). In individual TEM slices, SV density within all three zones is significantly reduced in LGC1fs mutants versus controls, with the difference increasing as the AZ membrane is approached (Fig. 8F). In the closest zone 1, the control density is 665±56 SV/µm2 (mean±s.e.m.), compared to 468±28 SV/µm2 in the LGC1-null mutants (Fig. 8F). Taken together with the above results, we suggest LGC1 loss elevates synaptic function by leading to greater SV release at an increased number of synapses. Increased neurotransmission translates to accelerated movement, causing rapid peristaltic waves and faster coordinated motion. However, strengthened NMJs also exhibit an impairment in replenishing activity-depleted SV pools in mutants. Thus, LGC1 regulates movement and NMJ synaptic function via activity-dependent changes in both SV cycling dynamics and SV release probability.

Anterograde and retrograde signals secreted into the NMJ synaptomatrix by both motor neuron and muscle cells orchestrate intercellular communication at the synapse (Rushton et al., 2020). The signaling ligands, their receptors and other molecules in the synaptomatrix environment are all very highly glycosylated, and mutation of many genes affecting glycosylation has strong effects on both coordinated movement and NMJ synaptic function (Broadie et al., 2011; Dani et al., 2012, 2014; Jumbo-Lucioni et al., 2014, 2016; Parkinson et al., 2013, 2016). Glycan-binding lectins interact with the NMJ synaptomatrix at many levels, to effectively modulate NMJ neurotransmission strength. In the extracellular space, lectins serve to regulate the distribution of glycans (as scaffolds, co-receptors and signaling platforms), thereby directing secreted and transmembrane protein function (Del Fresno et al., 2018; Sych et al., 2018; Choi et al., 2019). For example, we have extensively characterized the secreted C-type lectin Mind-the-Gap (MTG) at the Drosophila NMJ, which regulates glycan distribution, secreted trans-synaptic signaling ligands, synaptic position-specific (PS) integrins and the postsynaptic glutamate receptor (GluR) field, as well as the downstream internal Drosophila Pix–Pak–Dock signaling cascade and the critical DLG synaptic scaffold (Rohrbough et al., 2007; Rushton et al., 2009, 2012; Rohrbough and Broadie, 2010). These secreted MTG C-type lectin roles controlling postsynaptic function suggest that similar C-type lectin roles may govern presynaptic function.

In a previous systematic transgenic RNAi screen for NMJ modulator genes, we did indeed uncover a new secreted C-type lectin that strongly controls neurotransmission strength (Dani et al., 2012). Lectin-galC1 (LGC1) is an ideal candidate for modulation of synaptic function due to its characterized Ca2+-binding (C-type) and glycan-binding properties as a secreted protein (Tanji et al., 2006; Haq et al., 1996). To create a LGC1-specific null mutation, we here used CRISPR/Cas9 homology-directed DNA repair to create a targeted frame-shift mutation (Fig. 1A–C; Gratz et al., 2013). We paired this LGC1fs null mutant with a deletion removing the C-type lectin gene cluster containing LGC1 to characterize a new specific LGC1 antibody and to show that LGC1 is locally secreted at the NMJ (Fig. 1D–J). We then moved forward to study the effects of LGC1 loss on coordinated movement, which intimately depends on both NMJ synaptic architecture and function. Consistent with our initial discovery of increased NMJ neurotransmission following LGC1 RNAi knockdown (Dani et al., 2012), and earlier studies demonstrating LGC1 Ca2+-dependent binding functions (Tanji et al., 2006; Haq et al., 1996), we report here that LGC1-null mutants exhibit accelerated movement (Fig. 2). LGC1 loss increases both the rate of locomotion and the speed to complete complex coordinated behaviors when challenged. These findings suggest LGC1 acting locally at the NMJ is important for modulating locomotion and, particularly, complex coordinated motor function.

LGC1 is secreted at the NMJ, as demonstrated using non-permeabilized labeling (Rushton et al., 2012; Kopke et al., 2017), and concentrated within the extracellular synaptomatrix surrounding synaptic boutons; a space characterized by striking glycan accumulation (Parkinson et al., 2013; Rushton et al., 2020). LGC1-null mutants have absolutely no alterations in NMJ structure (Fig. 3A–F). This is surprising, since NMJ overelaboration is a common feature of glycosylation mutations (Rushton et al., 2020). For example, mutations in the key phosphomannomutase type 2 (PMM2) glycosylation enzyme strongly alter synaptomatrix pauci-mannose glycans to cause significant NMJ architectural expansion, including more synaptic branching and supernumerary boutons (Parkinson et al., 2016). Similarly, loss of UDP-sugar precursors, which are required for synaptic glycosylation also results in obvious NMJ structural overelaboration, with both increased branching and synaptic bouton number (Jumbo-Lucioni et al., 2014, 2016). However, synaptic structure and function are independently regulated, with separable glycan roles (Collins and DiAntonio, 2007; Miller et al., 2012). Consistently, LGC1-null mutants exhibit an elevated number of synapses within otherwise normal NMJ boutons, thereby exhibiting greater synaptic density (Fig. 3G–M). LGC1 nulls have more presynaptic BRP-positive active zones (AZs) and more apposed postsynaptic GluR fields, suggesting that a higher level of synaptic connectivity contributes to elevated neurotransmission strength.

LGC1 limits NMJ function by reducing SV fusion and evoked transmission (Fig. 4). An elevated LGC1 EJC amplitude reflects an increase in number of released vesicles (quantal content; Perry et al., 2020) in a more mobile readily releasable pool (RRP; Vaden et al., 2019). Since spontaneous mEJC amplitude does not change in LGC1 nulls (Fig. 4), an increase in postsynaptic GluR number or function is ruled out as a contributing factor (Bykhovskaia and Vasin, 2017; Kaeser and Regehr, 2017). In addition to increased synapse density, we hypothesized that there would be higher SV fusion probability from altered SV cycling in LGC1 mutants. FM1-43 dye imaging (Rodesch and Broadie, 2000; Kopke and Broadie, 2018) shows that LGC1 loss results in a smaller cycling SV pool with higher cycling rate (Fig. 5). A smaller SV pool could negate the effects of increased synapse number and SV fusion probability. However, most SVs do not participate detectably in transmission cycling, and smaller SV pools can show high transmission due to increased synaptic density and SV cycling rates (Denker et al., 2011). Moreover, altered SV endocytosis–exocytosis cycling efficacy can cause altered synaptic depression/recovery with high-frequency stimulation (HFS; Stevens et al., 2012). Stimulation frequency can change SV cycling dynamics, with higher frequencies triggering differential SV release and recovery mechanisms (Maeno-Hikichi et al., 2011). Consequently, we assayed NMJ function (depression and recovery) during varying levels of frequency stimulation, to stress neurotransmission machinery, and thus gain insight into the LGC1 mutant SV cycling mechanisms.

In LGC1 nulls, HFS causes more rapid fatigue and slower recovery (Fig. 6). Consistent with this, smaller SV cycling pools with higher turnover rates can increase basal transmission strength while impairing transmission maintenance during HFS demand (Baldelli et al., 2007; Doussau et al., 2017). Thus, the smaller cycling SV pool in LGC1 mutants (Fig. 5) can maintain elevated basal transmission while exhibiting impaired replenishment of SV pools and maintenance of transmission during HFS (Fig. 6). Other glutamatergic synapse classes maintain a small cycling RRP for high probability SV release during basal stimulation conditions, yet are unable to maintain transmission under conditions of greater demand (Xu-Friedman et al., 2001; Valera et al., 2012). Likewise, LGC1 nulls exhibit reduced membrane-associated SV pools at both the light microscope (Fig. 7) and electron microscope (Fig. 8) levels, including lower SV density at AZs. Other Drosophila mutants with increased NMJ neurotransmission can similarly show presynaptic SV depletion in TEM ultrastructural studies (Jia et al., 1993). At the AZ, SVs near the presynaptic membrane make up the RRP and linked cycling pools (Moulder and Mennerick, 2005), but internal SV pools can be held in ‘reserve’ or serve other functions apart from neurotransmitter release (Denker et al., 2011). Overall, our findings are consistent with LGC1 mutants exhibiting a smaller and faster SV cycling pool, with altered SV cycle dynamics during and following HFS stress.

Similar to what is seen with LGC1 mutants, the loss of a number of C-type lectin domain (CTLD) proteins causes synaptic defects. For example, Mincle-deficient mice suppress TNF induction, which lowers glutamate release from central synapses, resulting in a direct modulation of neurotransmission strength (Richardson and Williams, 2014; Ishikawa et al., 2019). Although LGC1 appears to have the opposite effect, the Drosophila TNF homolog (Eiger) downregulates excitatory amino acid transporters 1 and 2 (EAAT1/2; Igaki et al., 2009), and the EAAT1 mutation impairs locomotory peristaltic movement (Stacey et al., 2010). We show here that LGC1 limits locomotory peristaltic movement, suggesting a possible relationship between LGC1, Eiger and EAAT1 may be a useful avenue to explore in future studies. Another CTLD protein is Caenorhabditis elegans CLEC-38, which has been shown to regulate presynaptic organization and mediate changes in the integral SV protein Synaptobrevin (Kulkarni et al., 2008). As the vesicle SNARE (V-SNARE), Synaptobrevin drives SV fusion with the presynaptic membrane AZ to directly mediate neurotransmitter release (Haberman et al., 2012; Estes et al., 2000). Since LGC1 mutants display neurotransmission defects comparable to CLEC-38 mutants, such as similar increased SV fusion and impaired SV cycling, the exploration of downstream LGC1 interactions and Synaptobrevin regulation in the presynaptic terminal could offer more insight into the mechanisms by which LGC1 produces neurotransmission changes.

In conclusion, this study provides insight into the roles of a novel C-type lectin at a glutamatergic neuromuscular synapse. Beyond the previously characterized roles in development and immunity (Tanji et al., 2006; Haq et al., 1996), we find LGC1 secreted at the NMJ synapse modulates presynaptic function. Along with acting to limit synapse number and neurotransmission strength, LGC1 also regulates SV cycling to increase SV availability during basal levels of use, but decrease SV availability during high levels of use. In resting basal conditions, LGC1 caps NMJ neurotransmission strength to limit coordinated motor function. With intense stimulation, LGC1 loss causes more fatigue and decreases the recovery rate, indicating impaired activity-dependent SV cycling. We reveal functional defects with electrophysiology, and visualize SVs with both FM1-43 dye and TEM electron microscopy imaging. In future studies, we plan to dissect LGC1 mechanisms by focusing on SV pools and trafficking, as well as testing the possible involvement in signaling via Eiger, EAAT1 and downstream SV Synaptobrevin changes. Our work indicates that LGC1-dependent signaling fine-tunes the multiple presynaptic pools that exchange SVs differentially based on activity levels and release demand to control SV release probability (Bui and Glavinović, 2014; Guo et al., 2015; Millar et al., 2002). Beyond this LGC1 mechanism, we hope this new model will allow further exploration of SV cycling-dependent NMJ function in neuromuscular disease states.

Drosophila genetics

All stocks were reared on standard cornmeal, molasses, yeast and agar food medium under standard conditions of constant 25°C with a 12 h light–12 h dark cycle. The genetic background control was w1118. The deficiency Df(2 L)Exel6043 (BSC#7525) removes the entire lectin-galC1 (LGC1) gene, as well as two downstream homologs and several upstream genes. A CRISPR/Cas9-generated LGC1 frame-shift mutant (LGC1fs) was made and confirmed (described below). Mutants were assayed as LGC1fs/LGC1fs and LGC1fs/Df(2L)Exel6043. A UAS-LGC1::GFP line was made with LGC1 cDNA attached to GFP coding sequence with a flexible linker. This construct was injected into y1w67c23; P{CaryP}attP1 (BDSC#8621) embryos to insert on chromosome 2 at cytological location 55C4 (Bloomington Drosophila Stock Center). The ubiquitous Gal4 driver used was for the daughterless (da) gene (P{Gal4-da.G32}UH1; UH1-Gal4; Wodarz et al., 1995).

CRISPR/Cas9 editing

Both mutagenesis and mutant line isolation were conducted as previously described (Shilts and Broadie, 2017). Briefly, a frame-shift (fs) null LGC1 allele (LGC1fs) was generated using CRISPR/Cas9 genome editing (Gratz et al., 2013). The guide RNA (gRNA) oligonucleotide sequence targeting the LGC1 gene was:

5′-CACTGAAATGCTGAAGCTTAGTTT-3′.

The gRNA was assembled in a pCFD3-dU6:3gRNA vector (Addgene #49410; Port et al., 2014). Purified constructs were injected (250 ng/μl) into w1118, vas-Cas9 Drosophila embryos (BestGene, Inc.). Injected embryos were then crossed into the homogenized genetic background. PCR sequencing was used on isolated mutant alleles to verify nucleotide sequence changes. One characterized frame-shift null allele (LGC1fs) was used in all the phenotypic studies. For phenotype comparisons, LGC1fs was crossed with the molecularly defined Df(2L)Exel6043 deficiency to generate LGC1fs/Df.

Antibody generation

A new LGC1 antibody was generated against amino acids 21–186 of LGC1 (ABClonal, Woburn, Massachusetts, USA). This amino acid sequence represents the entire LGC1 protein, excluding only the signal peptide. The specific sequence used was: REKFSIQVNEGNTFGALVKAEPFTKINDGYYFFGTESLNWYEAYEKCRELNSELVTFETDQEFDAVTAFLTANGSRLTYWTSGNDLAKTGSHRWFTNAQRISSLRWARNQPDNAGQKEHCIHLGYIYKDSRKFELNDRPCSQDPNSLFKYICEAPEMETISIVVWK.

The sequence was determined to be antigenic and display no cross-reactivity with the Drosophila LGC1 homologs (CG33532 and CG33533). Three antiserums (16, 17, 18) were recovered and affinity-purified from three separate injected rabbits. LGC1 Ab16 was preabsorbed against the LGC1fs null mutants prior to use in this study.

Western blotting

Western blotting was performed as previously described (Vita and Broadie, 2017). Briefly, two adult males at 1 day post-eclosion (dpe) were homogenized in 1× NuPAGE LDS sample buffer (45 μl, Invitrogen, NP007), heated at 95°C with 2-mercaptoethanol (2-ME; 5 μl, Sigma-Aldrich, M7154) for 10 min, and then centrifuged at 7800 g for 10 min. Supernatant (10 μl) was loaded onto 4–12% Bis-Tris gels (Invitrogen, NP0336) and run with NuPage 1× MES running buffer (Invitrogen, NP002). Samples were transferred to nitrocellulose membranes (PROTRAN, NBA085C001EA) in NuPage transfer buffer with 20% methanol at 30 mA. Membranes were labeled for protein content with Ponceau S stain (Sigma-Aldrich, P7170). Odyssey blocking buffer diluted in TBS-T (10 mM Tris-HCl pH 8, 150 mM NaCl and 0.05% Tween 20) was used to block membranes for 1 h at room temperature (RT). Samples were incubated in rabbit anti-LGC1 (1:5000) and goat anti-GFP (1:2500; ab6662; Abcam) for 2 h at RT. Membranes were subsequently washed with TBS-T and incubated in Alexa Fluor 680 donkey anti-rabbit-IgG (1:10,000; Jackson Laboratory) and Alexa Fluor 800 donkey anti-goat-IgG (1:10,000; Jackson Laboratory). Blots were imaged with an Odyssey CLx infrared scanner (Licor).

Immunocytochemistry

Immunolabeling of wandering third-instar larvae was performed as previously described (Staples and Broadie, 2013). Briefly, staged animals were dissected in physiological saline (128 mM NaCl, 2 mM KCl, 4 mM MgCl2, 0.2 mM CaCl2, 70 mM sucrose, 5.5 mM trehalose, 5 mM HEPES, pH 7.2). A lateral cut was made at the posterior and extended anteriorly in a dorsal longitudinal incision, and the body walls were pinned down. The internal organs were removed, revealing body wall muscles, central nervous system (CNS) and motor nerves. The preparations were fixed in 4% paraformaldehyde (PFA) for 10 min, and then blocked with 0.5% bovine serum albumin (BSA) for 10 min. Preparations were either permeabilized with 0.2% Triton X-100 (for HRP, DLG, BRP, GluRIIC and VGLUT), or processed without detergent for extracellular only labeling (LGC1). Primary antibodies used were: rabbit anti-LGC1 (1:50; this study), mouse anti-Discs Large [DLG; 1:250; Developmental Studies Hybridoma Bank (DSHB), 4F3], goat anti-horse radish peroxidase (HRP) Alexa Fluor647-conjugated (1:300; Jackson Laboratories, 123-605-021), mouse anti-Bruchpilot (BRP; 1:100; DSHB, nc82), rabbit anti-glutamate receptor IIC (GluRIIC; 1:5000, Marrus and DiAntonio, 2004), and rabbit anti-vesicular glutamate receptor (VGLUT; 1:10,000; Daniels et al., 2004). Preparations were incubated in primary antibody overnight at 4°C. Secondary antibodies used were: goat anti-rabbit-IgG Alexa Fluor 488-conjugated, goat anti-mouse-IgG Alexa Fluor 555-conjugated, goat anti-mouse-IgG Alexa Fluor488-conjugated, and donkey anti-rabbit-IgG Alexa Fluor 555-conjugated, all at 1:300 (Invitrogen). Preparations were incubated in secondary antibodies for 2 h at RT. Preparations were mounted in Fluoromount G (Electron Microscopy Services) for confocal imaging.

Confocal imaging

Laser-scanning confocal microscopy was used for all imaging as previously described (Sears et al., 2019). Briefly, 40× water-immersion (for FM1-43) and 63× oil-immersion (for all immunocytochemistry) Plan Apo objective lens were used for imaging on a Zeiss LSM 510 META microscope. Muscle 4 NMJs in abdominal segments 3 and 4 were imaged in Z-stacks. All fluorescence quantifications were undertaken using ImageJ software (NIH, open source). Maximum intensity projections were made using the ‘Z Project Function’. Regions of interest (ROIs) were determined at the NMJ with the anti-HRP:633 channel using the ‘Threshold Function’. ROIs were subsequently applied to other channels, with the ‘Measure Function’ used to obtain average fluorescence intensities. The gray value intensity line graphs were created using the ‘Plot Profile Function’. BRP puncta were quantified using the ‘Find Maxima Function’. For NMJ structural analyses, type Ib synaptic boutons were defined as HRP- and DLG-positive varicosities with a diameter ≥2 μm. Branches were defined as containing ≥2 boutons.

Behavioral assays

Peristaltic contraction was quantified as reported previously (Shilts and Broadie, 2017). Briefly, an individual wandering third-instar larva was placed at the center of a 2% agarose plate and acclimated for 30 s at RT. With even illumination, peristaltic waves were video recorded using an Olympus dissection microscope and mounted Canon Rebel DLSR camera. Uninterrupted movement (1 min) was recorded for each animal. Videos were analyzed at 0.5× speed on Windows MPC-HC. Peristaltic motion was defined as the complete movement of a muscular contraction wave from posterior to anterior. Five complete waves were measured and averaged per animal for a single data point (N=1). Coordinated movement roll-over assays were performed as previously described (Jumbo-Lucioni et al., 2016). Briefly, an individual wandering third-instar larva was placed on a 2% agarose plate and acclimated for 30 s at RT. A fine paintbrush was used to gently roll the larva until the ventral midline was exactly upwards (t=0). A stopwatch was then used to record the time until the ventral middle was exactly downwards. The assay was repeated three times on the same animal, and the three times were then averaged for each single data point (N=1).

Electrophysiological recording

The two-electrode voltage-clamp (TEVC) electrophysiology configuration was employed as previously described (Kopke et al., 2020). Briefly, wandering third-instar larvae were dissected along the dorsal midline and the walls glued down (3M Vetbond Tissue Adhesive). The internal organs were removed to reveal the neuromusculature. The peripheral nerves were cut at the CNS exit points. Recordings were carried out at 18°C in physiological saline (see above). A 40× water-immersion objective lens on a Zeiss Axioskop microscope was used to image the preparation. Two microelectrodes (1-mm outer diameter borosilicate capillaries; World Precision Instruments) were inserted into muscle 6 of abdominal segments 3 or 4. Microelectrodes were ∼15 MΩ resistance when filled with 3 M KCl. The muscle was voltage-clamped at −60 mV (Axoclamp-2B amplifier). Spontaneous miniature excitatory junction current (mEJC) recordings were made in continuous 2 min sessions and then low-pass filtered. For nerve stimulation-evoked EJC recordings, a fire-polished glass suction electrode was used to suck up the motor nerve. A Grass (S88) stimulator was used to stimulate the motor nerve with 0.5 ms suprathreshold voltage stimuli at 0.2 Hz. EJC recordings were filtered at 2 kHz. To quantify EJC amplitude, 10 consecutive traces were averaged and the average peak value recorded. For high-frequency stimulation (HFS) train assays, motor nerves were stimulated at 5 Hz, 20 Hz or 50 Hz for 5 min. For recovery assays, ten consecutive EJC traces with 0.2 Hz stimulation were averaged at 30 s intervals following the HFS train. Clampex 9.0 was used for data acquisition, and Clampfit 9 was used for data analyses.

FM1-43 imaging

FM1-43 dye labeling was done as previously described (Kopke and Broadie, 2018). Briefly, staged wandering third-instar larvae were dissected in physiological saline (see above) and NMJ terminals labeled with anti-HRP:647 (100 μl; Sigma) added to the bath. Peripheral nerves were cut at the CNS, and the CNS removed. The bath solution was then replaced with FM1-43 (4 μM; Invitrogen) in 1.0 mM [Ca2+] physiological saline. As above, a glass suction electrode was used to suck up the motor nerve for stimulation (20 Hz, 5 min). The bath solution was then replaced several times in quick succession with Ca2+-free saline to halt synaptic vesicle (SV) cycling. Z-stacks of the FM1-43-loaded NMJ were taken with the Zeiss LSM 510 confocal microscope using a 40× water immersion objective. Next, the Ca2+-free saline was replaced with 1.0 mM [Ca2+] saline (without FM1-43) and the same motor nerve stimulated (20 Hz, 1 min) to drive SV exocytosis and dye release. The bath solution was then replaced several times in quick succession with Ca2+-free saline to halt SV cycling. Z-stacks of the same NMJ were then taken. Images were quantified with ImageJ (NIH) by setting a threshold using the anti-HRP:633 channel to highlight the NMJ. The FM1-43 dye channel was then selected, and the ‘Measure Function’ used to assay fluorescence intensity.

Electron microscopy

Synaptic transmission electron microscopy (TEM) was undertaken as previously described (Kopke et al., 2017). Briefly, staged wandering third-instar larvae were dissected as above and fixed overnight at 4°C in 2.5% glutaraldehyde. A secondary fixation was then performed in 1% osmium tetroxide for 1 h at RT. The preparations were washed in 0.1 M sodium cacodylate buffer three times (10 min each), and then ddH2O three times (15 min each). En bloc 2% uranyl acetate staining was undertaken for 2 h at RT in the dark. Preparations were again rinsed in ddH2O three times (15 min each), followed by an ethanol dehydration series (30, 50, 70, 90, 95, 100, 100%), propylene oxide infiltration, and resin embedding (Embed-812). Muscles 6 or 7 from abdominal segments 2 or 3 were dissected free and embedded in a semi-hardened resin block. Isolated muscles from four animals were put in each block, with three blocks for each genotype. Blocks were polymerized at 60°C for 48 h. Thick sections (1 µm) were cut from the blocks using a glass knife, stained with Toluidine Blue (Sigma-Aldrich) for 1 min on a Thermostat slide warmer (45°C), and then imaged on a compound microscope with a 100× objective (total magnification 1000×) for NMJ bouton identification. Once a bouton was found, ultrathin (50 nm) sections were cut using a DiATOME diamond knife on a Leica Ultracut UCT ultramicrotome, and then collected on uncoated 200 mesh copper grids. All TEM imaging was performed on a Philips CM10 transmission electron microscope operating at 80 kV, with images collected using a 4-megapixel AMT CCD camera. Bouton area, SSR area and SV size, number and distribution were all measured using ImageJ.

Statistical analyses

All study statistics were performed using Prism software (GraphPad, San Diego, CA). All data sets were subjected to Shapiro–Wilks normality tests. All comparisons between any two experimental groups were performed by using the two-tailed Welch's unequal variances t-test, or the Student's two-tailed unpaired t-test, as indicated. All comparisons between ≥3 experimental groups were performed by using one-way ANOVA tests, except when the paired data were assayed, for which two-way ANOVA tests were used, as indicated. All display graphs were made using Prism software to show all the individual data points (N indicated) as well as the mean±s.e.m. In all of the figures, significance is indicated as *P≤0.05, **P≤0.01, ***P≤0.001, ****P<0.0001 and P>0.05 (not significant, ns).

We are grateful to Daniel Okoye and Darius Booth for their technical input on this study. We thank both the Bloomington Drosophila Stock Center (BDSC; Indiana University, Bloomington, IN, USA) and the Vienna Drosophila Resource Center (VDRC; Vienna, Austria) for providing essential genetic lines for this study. We thank the Developmental Studies Hybridoma Bank (DSHB; University of Iowa, Iowa City, IA, USA) for providing essential antibodies. We are especially grateful to Dr Aaron DiAntonio (Washington University, St Louis, MO, USA) for the gift of both GluRIIC and VGLUT antibodies.

Author contributions

Conceptualization: M.B., K.B.; Methodology: M.B., E.R., D.L.K., K.B.; Software: M.B.; Validation: M.B.; Formal analysis: M.B.; Investigation: M.B., D.L.K.; Resources: E.R., D.L.K., K.B.; Writing - original draft: M.B.; Writing - review & editing: M.B., E.R., K.B.; Visualization: M.B., K.B.; Supervision: E.R., K.B.; Project administration: K.B.; Funding acquisition: M.B., K.B.

Funding

This work was fully supported by National Institutes of Health grant MH096832 to K.B. M.B. was also supported by the Littlejohn Fellowship for undergraduate research at Vanderbilt University. Deposited in PMC for release after 12 months.

Ao
,
J.
,
Ling
,
E.
and
Yu
,
X.-Q.
(
2007
).
Drosophila C-type lectins enhance cellular encapsulation
.
Mol. Immun.
44
,
2541
-
2548
.
Astorga
,
C.
,
Jorquera
,
R.
,
Ramírez
,
M.
,
Kohler
,
A.
,
López
,
E.
,
Delgado
,
R.
,
Córdova
,
A.
,
Olguín
,
P.
and
Sierralta
,
J.
(
2016
).
Presynaptic DLG regulates synaptic function through the localization of voltage-activated Ca2+ Channels
.
Sci. Rep.
6
,
32132
.
Baldelli
,
P.
,
Fassio
,
A.
,
Valtorta
,
F.
and
Benfenati
,
F.
(
2007
).
Lack of Synapsin I reduces the readily releasable pool of synaptic vesicles at central inhibitory synapses
.
J. Neurosci.
27
,
13520
-
13531
.
Betz
,
W. J.
and
Bewick
,
G. S.
(
1992
).
Optical analysis of synaptic vesicle recycling at the frog neuromuscular junction
.
Science.
255
,
200
-
203
.
Bodily
,
K. D.
,
Morrison
,
C. M.
,
Renden
,
R. B.
and
Broadie
,
K.
(
2001
).
A novel member of the Ig superfamily, Turtle, is a CNS-specific protein required for coordinated motor control
.
J. Neurosci.
21
,
3113
-
3125
.
Brand
,
A. H.
and
Perrimon
,
N.
(
1993
).
Targeted gene expression as a means of altering cell fates and generating dominant phenotypes
.
Development.
118
,
401
-
415
.
Broadie
,
K.
,
Baumgartner
,
S.
and
Prokop
,
A.
(
2011
).
Extracellular matrix and its receptors in Drosophila neural development
.
Dev. Neurobiol.
71
,
1102
-
1130
.
Bui
,
L.
and
Glavinović
,
M. I.
(
2014
).
Is replenishment of the readily releasable pool associated with vesicular movement?
Cogn. Neurodyn.
8
,
99
-
110
.
Bykhovskaia
,
M.
and
Vasin
,
A.
(
2017
).
Electrophysiological analysis of synaptic transmission in Drosophila
.
Dev. Biol.
6
,
e277
.
Chanaday
,
N. L.
,
Cousin
,
M. A.
,
Milosevic
,
I.
,
Watanabe
,
S.
and
Morgan
,
J. R.
(
2019
).
The synaptic vesicle cycle revisited: new insights into the modes and mechanisms
.
J. Neurosci.
39
,
8209
-
8216
.
Chandel
,
I.
,
Baker
,
R.
,
Nakamura
,
N.
and
Panin
,
V.
(
2019
).
Live imaging and analysis of muscle contractions in Drosophila embryo
.
J. Vis. Exp.
149
,
59404
.
Choi
,
H.-K.
,
Lee
,
D.
,
Singla
,
A.
,
Kwon
,
J. S.-I.
and
Wu
,
H.-J.
(
2019
).
The influence of heteromultivalency on lectin-glycan binding behavior
.
Glycobiology.
29
,
397
-
408
.
Collins
,
C. A.
and
DiAntonio
,
A.
(
2007
).
Synaptic development: insights from Drosophila
.
Curr. Opin. Neurobiol.
17
,
35
-
42
.
Dani
,
N.
and
Broadie
,
K.
(
2012
).
Glycosylated synaptomatrix regulation of trans-synaptic signaling
.
Dev. Neurobiol.
72
,
2
-
21
.
Dani
,
N.
,
Nahm
,
M.
,
Lee
,
S.
and
Broadie
,
K.
(
2012
).
A targeted glycan-related gene screen reveals heparan sulfate proteoglycan sulfation regulates WNT and BMP trans-synaptic signaling
.
PLoS Genet.
8
,
e1003031
.
Dani
,
N.
,
Zhu
,
H.
and
Broadie
,
K.
(
2014
).
Two protein N-acetylgalactosaminyl transferases regulate synaptic plasticity by activity-dependent regulation of integrin signaling
.
J. Neurosci.
34
,
13047
-
13065
.
Daniels
,
R. W.
,
Collins
,
C. A.
,
Gelfand
,
M. V.
,
Dant
,
J.
,
Brooks
,
E. S.
,
Krantz
,
D. E.
and
DiAntonio
,
A.
(
2004
).
Increased expression of the Drosophila vesicular glutamate transporter leads to excess glutamate release and a compensatory decrease in quantal content
.
J. Neurosci.
24
,
10466
-
10474
.
Dear
,
M. L.
,
Dani
,
N.
,
Parkinson
,
W.
,
Zhou
,
S.
and
Broadie
,
K.
(
2016
).
Two classes of matrix metalloproteinases reciprocally regulate synaptogenesis
.
Development
143
,
75
-
87
.
Del Fresno
,
C.
,
Iborra
,
S.
,
Saz-Leal
,
P.
,
Martínez-López
,
M.
and
Sancho
,
D.
(
2018
).
Flexible signaling of Myeloid C-Type Lectin receptors in immunity and inflammation
.
Front. Immun.
9
,
804
.
Denker
,
A.
,
Bethani
,
I.
,
Kröhnert
,
K.
,
Körber
,
C.
,
Horstmann
,
H.
,
Wilhelm
,
B. G.
,
Barysch
,
S. V.
,
Kuner
,
T.
,
Neher
,
E.
and
Rizzoli
,
S. O.
(
2011
).
A small pool of vesicles maintains synaptic activity in vivo
.
Proc. Natl. Acad. Sci. USA
108
,
17177
-
17182
.
Dityatev
,
A.
,
Schachner
,
M.
and
Sonderegger
,
P.
(
2010
).
The dual role of the extracellular matrix in synaptic plasticity and homeostasis
.
Nat. Rev. Neurosci.
11
,
735
-
746
.
Dodd
,
R. B.
and
Drickamer
,
K.
(
2001
).
Lectin-like proteins in model organisms: implications for evolution of carbohydrate-binding activity
.
Glycobiology
11
,
71
-
79
.
Doll
,
C. A.
and
Broadie
,
K.
(
2014
).
Impaired activity-dependent neural circuit assembly and refinement in autism spectrum disorder genetic models
.
Front. Cell. Neurosci.
8
,
30
.
Doussau
,
F.
,
Schmidt
,
H.
,
Dorgans
,
K.
,
Valera
,
A. M.
,
Poulain
,
B.
and
Isope
,
P.
(
2017
).
Frequency-dependent mobilization of heterogeneous pools of synaptic vesicles shapes presynaptic plasticity
.
eLife
6
,
e28935
.
Estes
,
P. S.
,
Ho
,
G. L. Y.
,
Narayanan
,
R.
and
Ramaswami
,
M.
(
2000
).
Synaptic localization and restricted diffusion of a Drosophila neuronal Synaptobrevin - Green Fluorescent Protein chimera in vivo
.
J. Neurogenet.
13
,
233
-
255
.
Fergestad
,
T.
,
Davis
,
W. S.
and
Broadie
,
K.
(
1999
).
The Stoned proteins regulate synaptic vesicle recycling in the presynaptic terminal
.
J. Neurosci.
19
,
5847
-
5860
.
Frappaolo
,
A.
,
Sechi
,
S.
,
Kumagai
,
T.
,
Karimpour-Ghahnavieh
,
A.
,
Tiemeyer
,
M.
and
Giansanti
,
M. G.
(
2018
).
Modeling congenital disorders of N-linked glycoprotein glycosylation in Drosophila melanogaster
.
Front. Genet.
9
,
436
.
Friedman
,
S. H.
,
Dani
,
N.
,
Rushton
,
E.
and
Broadie
,
K.
(
2013
).
Fragile X Mental Retardation Protein regulates trans-synaptic signaling in Drosophila
.
Dis. Models Mech.
6
,
1400
-
1413
.
Fushiki
,
A.
,
Zwart
,
M. F.
,
Kohsaka
,
H.
,
Fetter
,
R. D.
,
Cardona
,
A.
and
Nose
,
A.
(
2016
).
A circuit mechanism for the propagation of waves of muscle contraction in Drosophila
.
eLife
5
,
e13253
.
Gan
,
Q.
and
Watanabe
,
S.
(
2018
).
Synaptic vesicle endocytosis in different model systems
.
Front. Cell. Neurosci.
12
,
171
.
Gerkin
,
R. C.
,
Nauen
,
D. W.
,
Xu
,
F.
and
Bi
,
G.-Q.
(
2013
).
Homeostatic regulation of spontaneous and evoked synaptic transmission in two steps
.
Mol. Brain
6
,
38
.
Gjorgjieva
,
J.
,
Berni
,
J.
,
Evers
,
J. F.
and
Eglen
,
S. J.
(
2013
).
Neural circuits for peristaltic wave propagation in crawling Drosophila larvae: analysis and modeling
.
Front. Comput. Neurosci.
7
,
24
.
Gokcezade
,
J.
,
Sienski
,
G.
and
Duchek
,
P.
(
2014
).
Efficient CRISPR/Cas9 plasmids for rapid and versatile genome editing in Drosophila
.
G3 (Bethesda)
4
,
2279
-
2282
.
Gratz
,
S. J.
,
Wildonger
,
J.
,
Harrison
,
M. M.
and
O'Connor-Giles
,
K. M.
(
2013
).
CRISPR/Cas9-mediated genome engineering and the promise of designer flies on demand
.
Fly
7
,
249
-
255
.
Grillner
,
S.
(
2003
).
The motor infrastructure: from ion channels to neuronal networks
.
Nat. Rev. Neurosci.
4
,
573
-
586
.
Guo
,
J.
,
Ge
,
J.-L.
,
Hao
,
M.
,
Sun
,
Z.-C.
,
Wu
,
X.-S.
,
Zhu
,
J.-B.
,
Wang
,
W.
,
Yao
,
P.-T.
,
Lin
,
W.
and
Xue
,
L.
(
2015
).
A three-pool model dissecting readily releasable pool replenishment at the Calyx of Held
.
Sci. Rep.
5
,
9517
.
Haberman
,
A.
,
Williamson
,
W. R.
,
Epstein
,
D.
,
Wang
,
D.
,
Rina
,
S.
,
Meinertzhagen
,
I. A.
and
Hiesinger
,
P. R.
(
2012
).
The synaptic vesicle SNARE neuronal Synaptobrevin promotes endolysosomal degradation and prevents neurodegeneration
.
J. Cell Biol.
196
,
261
-
276
.
Haq
,
S.
,
Kubo
,
T.
,
Kurata
,
S.
,
Kobayashi
,
A.
and
Natori
,
S.
(
1996
).
Purification, characterization, and cDNA cloning of a galactose-specific C-type lectin from Drosophila melanogaster
.
J. Biol. Chem.
271
,
20213
-
20218
.
Heigwer
,
F.
,
Port
,
F.
and
Boutros
,
M.
(
2018
).
RNA interference (RNAi) screening in Drosophila
.
Genetics
208
,
853
-
874
.
Igaki
,
T.
,
Pastor-Pareja
,
J. C.
,
Aonuma
,
H.
,
Miura
,
M.
and
Xu
,
T.
(
2009
).
Intrinsic tumor suppression and epithelial maintenance by endocytic activation of Eiger/TNF signaling in Drosophila
.
Dev. Cell
16
,
458
-
465
.
Ishikawa
,
A.
,
Miyake
,
Y.
,
Kobayashi
,
K.
,
Murata
,
Y.
,
Iizasa
,
S.
,
Iizasa
,
E.
,
Yamasaki
,
S.
,
Hirakawa
,
N.
,
Hara
,
H.
,
Yoshida
,
H.
et al. 
(
2019
).
Essential roles of C-type lectin Mincle in induction of neuropathic pain in mice
.
Sci. Rep.
9
,
872
.
Jia
,
X.-X.
,
Gorczyca
,
M.
and
Budnik
,
V.
(
1993
).
Ultrastructure of neuromuscular junctions in Drosophila: comparison of wild type and mutants with increased excitability
.
J. Neurobiol.
24
,
1025
-
1044
.
Jumbo-Lucioni
,
P.
,
Parkinson
,
W.
and
Broadie
,
K.
(
2014
).
Overelaborated synaptic architecture and reduced synaptomatrix glycosylation in a Drosophila classic galactosemia disease model
.
Dis. Model Mech.
7
,
1365
-
1378
.
Jumbo-Lucioni
,
P. P.
,
Parkinson
,
W. M.
,
Kopke
,
D. L.
and
Broadie
,
K.
(
2016
).
Coordinated movement, neuromuscular synaptogenesis and trans-synaptic signaling defects in Drosophila galactosemia models
.
Hum. Mol. Gen.
25
,
3699
-
3714
.
Kaeser
,
P. S.
and
Regehr
,
W. G.
(
2017
).
The readily releasable pool of synaptic vesicles
.
Curr. Opin. Neurobiol.
43
,
63
-
70
.
Kamimura
,
K.
and
Maeda
,
N.
(
2017
).
Heparan sulfate proteoglycans in Drosophila neuromuscular development
.
Biochim. Biophys. Acta Gen. Subj.
1861
,
2442
-
2446
.
Kauwe
,
G.
and
Isacoff
,
E. Y.
(
2013
).
Rapid feedback regulation of synaptic efficacy during high-frequency activity at the Drosophila larval neuromuscular junction
.
Proc. Natl. Acad. Sci. USA
110
,
9142
-
9147
.
Keller
,
B. G.
and
Rademacher
,
C.
(
2019
).
Allostery in C-type lectins
.
Curr. Opin. Struct. Biol.
62
,
31
-
38
.
Kittel
,
R. J.
,
Wichmann
,
C.
,
Rasse
,
T. M.
,
Fouquet
,
W.
,
Schmidt
,
M.
,
Schmid
,
A.
,
Wagh
,
D. A.
,
Pawlu
,
C.
,
Kellner
,
R. R.
,
Willig
,
K. I.
et al. 
(
2006
).
Bruchpilot promotes active zone assembly, Ca2+ channel clustering, and vesicle release
.
Science.
312
,
1051
-
1054
.
Kopke
,
D. L.
and
Broadie
,
K.
(
2018
).
FM dye cycling at the synapse: comparing high potassium depolarization, electrical and channelrhodopsin stimulation
.
J. Vis. Exp.
135
,
57765
.
Kopke
,
D. L.
,
Lima
,
S. C.
,
Alexandre
,
C.
and
Broadie
,
K.
(
2017
).
Notum coordinates synapse development via extracellular regulation of Wingless trans-synaptic signaling
.
Development
144
,
3499
-
3510
.
Kopke
,
D. L.
,
Leahy
,
S. N.
,
Vita
,
D. J.
,
Lima
,
S. C.
,
Newman
,
Z. L.
and
Broadie
,
K.
(
2020
).
Carrier of Wingless (Cow) regulation of Drosophila neuromuscular junction development
.
eNeuro
7
,
ENEURO.0285-19.2020
.
Kulkarni
,
G.
,
Li
,
H.
and
Wadsworth
,
W. G.
(
2008
).
CLEC-38, a transmembrane protein with C-type lectin-like domains, negatively regulates UNC-40-mediated axon outgrowth and promotes presynaptic development in Caenorhabditis elegans
.
J. Neurosci.
28
,
4541
-
4550
.
Maeno-Hikichi
,
Y.
,
Polo-Parada
,
L.
,
Kastanenka
,
K. V.
and
Landmesser
,
L. T.
(
2011
).
Frequency-dependent modes of synaptic vesicle endocytosis and exocytosis at adult mouse neuromuscular junctions
.
J. Neurosci.
31
,
1093
-
1105
.
Marrus
,
S. B.
and
DiAntonio
,
A.
(
2004
).
Preferential localization of glutamate receptors opposite sites of high presynaptic release
.
Curr. Biol.
14
,
924
-
931
.
Marrus
,
S. B.
,
Portman
,
S. J.
,
Allen
,
M. J.
,
Moffat
,
K. G.
and
DiAntonio
,
A.
(
2004
).
Differential localization of glutamate receptor subunits at the Drosophila neuromuscular junction
.
J. Neurosci.
24
,
1406
-
1415
.
Martin
,
C. A.
and
Krantz
,
D. E.
(
2014
).
Drosophila melanogaster as a genetic model system to study neurotransmitter transporters
.
Neurochem. Internat.
73
,
71
-
88
.
Menon
,
K. P.
,
Carrillo
,
R. A.
and
Zinn
,
K.
(
2013
).
Development and plasticity of the Drosophila larval neuromuscular junction
.
Dev. Biol.
2
,
647
-
670
.
Millar
,
A. G.
,
Bradacs
,
H.
,
Charlton
,
M. P.
and
Atwood
,
H. L.
(
2002
).
Inverse relationship between release probability and readily releasable vesicles in depressing and facilitating synapses
.
J. Neurosci.
22
,
9661
-
9667
.
Miller
,
D. L.
,
Ballard
,
S. L.
and
Ganetzky
,
B.
(
2012
).
Analysis of synaptic growth and function in Drosophila with an extended larval stage
.
J. Neurosci.
32
,
13776
-
13786
.
Moulder
,
K. L.
and
Mennerick
,
S.
(
2005
).
Reluctant vesicles contribute to the total readily releasable pool in glutamatergic hippocampal neurons
.
J. Neurosci.
25
,
3842
-
3850
.
Newman
,
Z. L.
,
Hoagland
,
A.
,
Aghi
,
K.
,
Worden
,
K.
,
Levy
,
S. L.
,
Son
,
J. H.
,
Lee
,
L. P.
and
Isacoff
,
E. Y.
(
2017
).
Input-specific plasticity and homeostasis at the Drosophila larval neuromuscular junction
.
Neuron
93
,
1388
-
1404
.
Nichols
,
C. D.
,
Becnel
,
J.
and
Pandey
,
U. B.
(
2012
).
Methods to assay Drosophila behavior
.
J. Vis. Exp.
61
,
3795
.
Orso
,
G.
,
Martinuzzi
,
A.
,
Rossetto
,
M. G.
,
Sartori
,
E.
,
Feany
,
M.
and
Daga
,
A.
(
2005
).
Disease-related phenotypes in a Drosophila model of hereditary spastic paraplegia are ameliorated by treatment with vinblastine
.
J. Clin. Invest.
115
,
3026
-
3034
.
Pan
,
B.
and
Zucker
,
R. S.
(
2009
).
A general model of synaptic transmission and short-term plasticity
.
Neuron
62
,
539
-
554
.
Parkinson
,
W. M.
,
Dear
,
M. L.
,
Rushton
,
E.
and
Broadie
,
K.
(
2013
).
N-glycosylation requirements in neuromuscular synaptogenesis
.
Development
140
,
4970
-
4981
.
Parkinson
,
W. M.
,
Dookwah
,
M.
,
Dear
,
M. L.
,
Gatto
,
C. L.
,
Aoki
,
K.
,
Tiemeyer
,
M.
and
Broadie
,
K.
(
2016
).
Synaptic roles for phosphomannomutase type 2 in a new Drosophila congenital disorder of glycosylation disease model
.
Dis. Models Mech.
9
,
513
-
527
.
Perry
,
S.
,
Goel
,
P.
,
Tran
,
N. L.
,
Pinales
,
C.
,
Buser
,
C.
,
Miller
,
D. L.
,
Ganetzky
,
B.
and
Dickman
,
D.
(
2020
).
Developmental arrest of Drosophila larvae elicits presynaptic depression and enables prolonged studies of neurodegeneration
.
Development
147
,
dev186312
.
Port
,
F.
,
Chen
,
H.
,
Lee
,
T.
and
Bullock
,
S. L.
(
2014
).
Optimized CRISPR/Cas tools for efficient germline and somatic genome engineering in Drosophila
.
Proc. Natl. Acad. Sci. USA
108
,
E2967
-
E2976
.
Pulver
,
S. R.
,
Bayley
,
T. G.
,
Taylor
,
A. L.
,
Berni
,
J.
,
Bate
,
M.
and
Hedwig
,
B.
(
2015
).
Imaging fictive locomotor patterns in larval Drosophila
.
J. Neurophys.
114
,
2564
-
2577
.
Rasse
,
T. M.
,
Fouquet
,
W.
,
Schmid
,
A.
,
Kittel
,
R. J.
,
Mertel
,
S.
,
Sigrist
,
C. B.
,
Schmidt
,
M.
,
Guzman
,
A.
,
Merino
,
C.
,
Qin
,
G.
et al. 
(
2005
).
Glutamate receptor dynamics organizing synapse formation in vivo
.
Nat. Neurosci.
8
,
898
-
905
.
Richardson
,
M. B.
and
Williams
,
S. J.
(
2014
).
MCL and Mincle: C-Type lectin receptors that sense damaged self and pathogen-associated molecular patterns
.
Front. Immun.
5
,
288
.
Rodesch
,
C. K.
and
Broadie
,
K.
(
2000
).
Genetic studies in Drosophila: vesicle pools and cytoskeleton-based regulation of synaptic transmission
.
Neuroreport
11
,
45
-
53
.
Rohrbough
,
J.
and
Broadie
,
K.
(
2010
).
Anterograde Jelly belly ligand to Alk receptor signaling at developing synapses is regulated by Mind the gap
.
Development
137
,
3523
-
3533
.
Rohrbough
,
J.
,
Pinto
,
S.
,
Mihalek
,
R. M.
,
Tully
,
T.
and
Broadie
,
K.
(
1999
).
Latheo, a Drosophila gene involved in learning, regulates functional synaptic plasticity
.
Neuron
23
,
55
-
70
.
Rohrbough
,
J.
,
Rushton
,
E.
,
Woodruff
,
E.
,
Fergestad
,
T.
,
Vigneswaran
,
K.
and
Broadie
,
K.
(
2007
).
Presynaptic establishment of the synaptic cleft extracellular matrix is required for post-synaptic differentiation
.
Genes Dev.
21
,
2607
-
2628
.
Rudling
,
J. E.
,
Drever
,
B. D.
,
Reid
,
B.
and
Bewick
,
G. S.
(
2018
).
Importance of full-collapse vesicle exocytosis for synaptic fatigue-resistance at rat fast and slow muscle neuromuscular junctions
.
Inter. J. Mol. Sci.
19
,
1936
.
Rushton
,
E.
,
Rohrbough
,
J.
and
Broadie
,
K.
(
2009
).
Presynaptic secretion of Mind-the-Gap organizes the synaptic extracellular matrix-integrin interface and postsynaptic environments
.
Dev. Dyn.
238
,
554
-
571
.
Rushton
,
E.
,
Rohrbough
,
J.
,
Deutsch
,
K.
and
Broadie
,
K.
(
2012
).
Structure-function analysis of endogenous lectin Mind-the-Gap in synaptogenesis
.
Dev. Neurobio.
72
,
1161
-
1179
.
Rushton
,
E.
,
Kopke
,
D. L.
and
Broadie
,
K.
(
2020
).
Extracellular heparan sulfate proteoglycans and glycan-binding lectins orchestrate trans-synaptic signaling
.
J. Cell Sci.
133
,
jcs244186
.
Schuster
,
C. M.
,
Ultsch
,
A.
,
Schloss
,
P.
,
Cox
,
J. A.
,
Schmitt
,
B.
and
Betz
,
H.
(
1991
).
Molecular cloning of an invertebrate glutamate receptor subunit expressed in Drosophila muscle
.
Science
254
,
112
-
114
.
Scott
,
H.
and
Panin
,
V. M.
(
2014
).
The role of protein N-glycosylation in neural transmission
.
Glycobiology
24
,
407
-
417
.
Sears
,
J. C.
,
Choi
,
W. J.
and
Broadie
,
K.
(
2019
).
Fragile X Mental Retardation Protein positively regulates PKA anchor Rugose and PKA activity to control actin assembly in learning/memory circuitry
.
Neurobiol. Dis.
127
,
53
-
64
.
Seinen
,
E.
,
Burgerhof
,
J. G. M.
,
Jansen
,
R. C.
and
Sibon
,
O. C. M.
(
2011
).
RNAi-induced off-target effects in Drosophila melanogaster: frequencies and solutions
.
Brief Funct. Genomics
10
,
206
-
214
.
Shilts
,
J.
and
Broadie
,
K.
(
2017
).
Secreted tissue inhibitor of matrix metalloproteinase restricts trans-synaptic signaling to coordinate synaptogenesis
.
J. Cell Sci.
130
,
2344
-
2358
.
Stacey
,
S. M.
,
Muraro
,
N. I.
,
Peco
,
E.
,
Labbé
,
A.
,
Thomas
,
G. B.
,
Baines
,
R. A.
and
van Meyel
,
D. J.
(
2010
).
Drosophila glial glutamate transporter Eaat1 is regulated by fringe-mediated notch signaling and is essential for larval locomotion
.
J. Neurosci.
30
,
14446
-
14457
.
Staples
,
J.
and
Broadie
,
K.
(
2013
).
The cell polarity scaffold Lethal Giant Larvae regulates synapse morphology and function
.
J. Cell Sci.
126
,
1992
-
2003
.
Stevens
,
R. J.
,
Akbergenova
,
Y.
,
Jorquera
,
R. A.
and
Littleton
,
J. T.
(
2012
).
Abnormal synaptic vesicle biogenesis in Drosophila synaptogyrin mutants
.
J. Neurosci.
32
,
18054
-
18067
.
Sun
,
S.-C.
and
Faye
,
I.
(
1992
).
Cecropia immunoresponsive factor, an insect immunoresponsive factor with DNA-binding properties similar to nuclear-factor kappa B
.
Eur. J. Biochem.
204
,
885
-
892
.
Sych
,
T.
,
Mély
,
Y.
and
Römer
,
W.
(
2018
).
Lipid self-assembly and lectin-induced reorganization of the plasma membrane
.
Philos. Trans. R. Soc.
373
,
20170117
.
Tanji
,
T.
,
Ohashi-Kobayashi
,
A.
and
Natori
,
S.
(
2006
).
Participation of a galactose-specific C-type lectin in Drosophila immunity
.
Biochem. J.
396
,
127
-
138
.
Theopold
,
U.
,
Rissler
,
M.
,
Fabbri
,
M.
,
Schmidt
,
O.
and
Natori
,
S.
(
1999
).
Insect glycobiology: a lectin multigene family in Drosophila melanogaster
.
Biochem. Biophys. Res. Commun.
261
,
923
-
927
.
Thomas
,
U.
,
Kim
,
E.
,
Kuhlendahl
,
S.
,
Koh
,
Y. H.
,
Gundelfinger
,
E. D.
,
Sheng
,
M.
,
Garner
,
C. C.
and
Budnik
,
V.
(
1997
).
Synaptic clustering of the cell adhesion molecule Fasciclin II by discs-large and its role in the regulation of presynaptic structure
.
Neuron.
91
,
787
-
799
.
Trier
,
N.
,
Hansen
,
P.
and
Houen
,
G.
(
2019
).
Peptides, antibodies, peptide antibodies and more
.
Inter. J. Mol. Sci.
20
,
6289
.
Vaden
,
J. H.
,
Banumurthy
,
G.
,
Gusarevich
,
E. S.
,
Overstreet-Wadiche
,
L.
and
Wadiche
,
J. I.
(
2019
).
The readily-releasable pool dynamically regulates multivesicular release
.
eLife.
8
,
e47434
.
Valera
,
A. M.
,
Doussau
,
F.
,
Poulain
,
B.
,
Barbour
,
B.
and
Isope
,
P.
(
2012
).
Adaptation of granule cell to Purkinje cell synapses to high-frequency transmission
.
J. Neurosci.
32
,
3267
-
3280
.
Vijayakrishnan
,
N.
,
Woodruff
,
E. A.
and
Broadie
,
K.
(
2009
).
Rolling Blackout is required for bulk endocytosis in non-neuronal cells and neuronal synapses
.
J. Cell Sci.
122
,
114
-
125
.
Vita
,
D. J.
and
Broadie
,
K.
(
2017
).
ESCRT-III membrane trafficking misregulation contributes to Fragile X syndrome synaptic defects
.
Sci. Rep.
7
,
8683
.
Wagh
,
D. A.
,
Rasse
,
T. M.
,
Asan
,
E.
,
Hofbauer
,
A.
,
Schwenkert
,
I.
,
Durrbeck
,
H.
,
Buchner
,
S.
,
Dabauvalle
,
M. C.
,
Schmidt
,
M.
,
Qin
,
G.
et al. 
(
2006
).
Bruchpilot, a protein with homology to ELKS/CAST, is required for structural integrity and function of synaptic active zones in Drosophila
.
Neuron.
49
,
833
-
844
.
Wagner
,
N.
(
2017
).
Ultrastructural comparison of the Drosophila larval and adult ventral abdominal neuromuscular junction
.
J. Morph.
278
,
987
-
996
.
Wilson
,
N. R.
,
Kang
,
J.
,
Hueske
,
E. V.
,
Leung
,
T.
,
Varoqui
,
H.
,
Murnick
,
J. G.
,
Erickson
,
J. D.
and
Liu
,
G.
(
2005
).
Presynaptic regulation of quantal size by the vesicular glutamate transporter VGLUT1
.
J. Neurosci.
25
,
6221
-
6234
.
Wodarz
,
A.
,
Hinz
,
U.
,
Engelbert
,
M.
and
Knust
,
E.
(
1995
).
Expression of crumbs confers apical character on plasma membrane domains of ectodermal epithelia of Drosophila
.
Cell
82
,
67
-
76
.
Xu-Friedman
,
M. A.
,
Harris
,
K. M.
and
Regehr
,
W. G.
(
2001
).
Three-dimensional comparison of ultrastructural characteristics at depressing and facilitating synapses onto cerebellar Purkinje cells
.
J. Neurosci.
21
,
6666
-
6672
.

Competing interests

The authors declare no competing or financial interests.